.‘ ~.. 1......» J. 2.. .3: :5}. .I. : s . ... p\.’ a; .1. Q- 03! , ‘ , m.A.I..OhlJA(.rd - i 5 v . v . V r ‘ ‘ I. 5 .f .. . £53 :3: . . . ,.I.~:,.l._15ur , V .v 1:5: ‘ .ar 1 r. (50. . o... .. .- 0: . v IIIIIIII'IIIII II IIIIIIIIIIIIIIIIIIIIII‘III '2 < 4 3 3 1293 00781 579 LIBRARY Michigan State University This is to certify that the dissertation entitled prepara‘hon And SPccf'rascoPir. Characfinzafioa O-F H 54.:- OXIdczfior) Siafe-S o-F Hemoprbfim Mack/5 presented by A3445! Sale/71. has been accepted towards fulfillment of the requirements for Ph'b ' degree in agar/5‘7"? ./ C I V k dag Major professor / Date @fi / 7,. 453 J PLACE It RETURN BOX to remove this Moat from your record. TO AVOID FINES return on or before die due DATE DUE DATE DUE DATE DUE "‘1 i I: 37F MSU is An Afflnnative AetioniEquul Opportunity inflation amm- PREPARATION AND SPECTROSCOPIC CHARACTERIZATION OF HIGHER OXIDATION STATES OF HEMOPROTEIN MODELS By Asaad Salehi A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1988 ‘4) 0 (\l 567 ABSTRACT PREPARATION AND SPECTROSCOPIC CHARACTERIZATION OF HIGHER OXIDATION STATES OF HEMOPROTEIN MODELS By Asaad Salehi This study utilizes synthetic octaethylporphyrin (OEP) complexes to establish vibrational, electronic and structural characteristics of hemoproteins' iron-porphyrin centers. The emphasis is primarily placed on the ring-centered n-cation radical state which may be involved in the reaction cycles of some hemoprotein enzymes and is also present in photosynthetic chlorophylls. A systematic study of these model radicals should aid detection and assignment of this particular mode of oxidation in the investigation of hemoprotein catalytic cycles by Resonance Raman spectroscopy. Metalloporphyrins (MP5) were oxidized by chemical methods using oxidants such as AgClO4, Fe(CIO4)3, Brz and pheno+-sbc16' in dichloromethane. The progress of the reaction was followed spectrophoto- metrically until complete oxidation was achieved. The color changes associated with radical formation are diagnostic: typically the initial pink or orange color of the neutral MP changes to green or black upon oxidation. The 363.8 nm Ar ion laser is the principal excitation source in our RR investigation as this wavelength is in resonance with the near-UV absorption bands of the radicals and, consequently, it serves to minimize contributions from the neutral porphyrin contaminants which absorb at longer wavelengths. The vibrational modes mainly in the 1350-1700 cm'1 region are Asaad Salehi examined owing to their sensitivity to the metal oxidation, spin and coordination states, and most importantly, they feature a linear correlation to the macrocycle core-size. Our RR data indicate that upon removal of an electron from the highest filled porphyrin n-bonding orbitals, alu or a2u, the frequencies of vibrations which are predominantly Cbe character increase dramatically (20- 30 cm’l); whereas, those with CaCm and CaN character decrease (3-10 cm'l) in the RR spectra of OEP radical complexes relative to the neutral precursors. These vibrational trends seem to hold in all the OEP complexes and is consistent with 2A1u radical symmetry assignment. This observation, however, contradicts the previous classification of CoIIIOEP+'2ClO4' as a 2A2u radical. Our data also suggest that the original classifica-tion of the radical optical spectra into two types, believed to arise from 2A1u and 2A2u ground states, may actually be caused by the planar vs. ruffled conformation of the macrocycle at the oxidized level. Metallochlorin (MC) n-cation radicals, on the other hand, feature a more complex vibrational spectra in the high-frequency region owing to the macrocycle reduced symmetry and increased conformational flexibility which result in more RR-allowed bands and increased likelihood of changes in mode composition, relative to their MP analogs. This is demonstrated by a vibrational mode in the 1500 cm"1 region which increases in frequency upon oxidation, in contrast with its MP counterpart. This characteristic is believed to result from both Cbe and CaCm contribution to this mode in MCs which differs from predominantly CaCm character of the corresponding mode in MP5. TO MY PARENTS ii ACKNOWLEDGEMENTS I wish to express my deep appreciation to my advisors, Drs. Chris K. Chang and Gerald T. Babcock for their guidance, encouragement and friendship during the course of this work. I am also grateful to Drs. Michael Rathke and Alexander Popov for serving on my committee. Financial support provided by the Department of Chemistry at Michigan State University is gratefully acknowledged. Many thanks go to my good friend and colleague Tony Oertling whose contribution to this study has been invaluable. I wish to thank all of the members of Dr. Chang's and Dr. Babcock's group for their help, friendship and support. I am especially grateful to Myoung S. Koo, Weishih Wu and Juan L.-Garriga whose friendship in and out of the lab will always be remembered; My special thanks go to a good friend, Mike Tornaritis, who has been like a brother to me during the past few years. I appreciate the help of Beverly Carnahan in typing my thesis and Ellen Rzepka for her sincerity and help in keeping my academic records accurate. Finally, I wish to thank my parents for their faith in me, their continuous love, encouragement and support. I dedicate this work to them. iii TABLE OF CONTENTS Page LIST OF TABLES ............................................................................................................ iv LIST OF FIGURES .......................................................................................................... vi Chapter 1 - INTRODUCTION Research Objectives ........................................................................................... 4 Porphyrin n-cation Radicals .............................................................................. 6 Metal-porphyrin Interactions ............................................................................ 9 Site-Selective Oxidation: Metal vs. Porphyrin Centered Oxidation Induction ..................................................................... 11 Porphyrin Peripheral Substituents ................................................................ 12 Metal Axial Ligand(s) ........................................................................................ 14 Solvent (or Medium) ........................................................................................ 16 Temperature ....................................................................................................... 17 Differention Between Metal vs. Porphyrin- Based Oxidation Reactions - Spectrosc0pic Methods ............................... 18 1. Optical Absorption Spectroscopy ............................................................... 18 2. ESR Method ................................................................................................... 23 3. Magnetic Susceptibility Measurement ..................................................... 26 4. Electrochemical Regularities ...................................................................... 28 Other Physicochemical Methods .................................................................... 29 5. Proton NMR of Metalloporphyrin n-cation Radicals ......................................................................................... 29 6. IR of Metalloporphyrin n-Cation Radical ................................................ 30 7. Resonance Raman ........................................................................................ 33 An Overview of this Thesis ............................................................................ 35 CHAPTER H - PREPARATION METHODS AND SPECTROSCOPIC PROPERTIES OF OXIDIZED COBALT PORPHYRIN S 1. Preparative Methods .................................................................................... 44 Instrumentation ............................................................................................ 47 2. Electronic Absorption Spectra .................................................................... 47 3. ESR And Magnetic Susceptibility .............................................................. 55 Infrared Diagnostic Band of the Cation Radicals ................................... 58 iv Page RR Vibrational Analysis of CoHOEP and its Oxidized Products-Spectral Characteristics of the Cation Radicals in the High-Frequency Region ............................ 60 CHAPTER III - IRON PROPHYRIN TI-CATION RADICALS: SOLUTION RESONANCE RAMAN SPECTRA OF OEP+°Fe(X)(X') Experimental Section ........................................................................................ 75 Materials and Methods ..................................................................................... 75 Chemical Ox1dations ....................... 76 Results and Discussion ..................................................................................... 76 High Frequency RR Scattering from Iron Porphyrin-IX n-cation Radical ....................................................................... 83 CHAPTER IV - RESONANCE RAMAN SPECTRA OF THE I'I-CATION RADICALS OF COPPER, COBALT, AND NICKEL METHYLOCTAETHYLCHLORINS: VIBRATIONAL CHARACTERISTICS OF CHLOROPHYLL MODELS Materials and Methods ..................................................................................... 92 Results and Discussion ..................................................................................... 92 Supplementary Material—Preparation of trans- octaethylchlorin—dzmfi),-d2(y,5) and -d4 as well as octaethylporphyrin—dz using acid-catalyzed exchange reaction ........................................................................................... 105 Experimental ..................................................................................................... 1 07 trans-octaethylchlorin-d2(y,5) ..................................................... 107 trans-06218thylchlorin-d4 ............................................................................... 110 trans-octaethylchlorin-d2(a,B) ...................................................................... 1 10 octaethylporphyrin-dz .................................................................................... 110 Conclusion ........................................................................................................ 11 1 CHAPTER V - NOVEL B-PYRROLIC SUBSTITUTION REACTIONS OF TETRAPERFLUOROPHENYL PORPHYRIN General Characteristics a. Reduced Affinity to Metal Ions ................................................................ 115 b. Resistance Toward Acid Demetallation ................................................. 115 Page c. Reduced Proton Affinity ............................................................................ 115 d. Optical Absorption Properties .................................................................. 116 e. 'H—N MR ........................................................................................................ 118 f. Reversible Oxidation Reactions of MTFPP Derivatives .................................................................................................. 119 Experimental ..................................................................................................... 120 Results ................................................................................................................ 121 Cu(II) and Ni(II) Adducts ................................................................ 125 A301) Adducts .................................................................................... 130 Zn (II) Adducts ................................................................................... 132 Discussion .......................................................................................................... 132 CHAPTER VI - METAL-AXIAL LIGAN D VIBRATIONS IN METALLOPORPHYRINS, THEIR IT-CATION RADICALS AND OTHER RELATED SYSTEMS Experimental ..................................................................................................... 142 Results and Discussion ................................................................................... 142 vi 1-1 1-2 1-3 5-1 5-2 LIST OF TABLES Page Spin Densities in Porphyrin Cation Radical ......................................... 8 The Nature of Axial Ligand(s) in Some Hemoproteins ................... 14 Proton and Deuteron NMR Chemical Shifts for Copper and Silver Porphyrins and Their One-Electron Oxidation Products ....................................................................................................... 31 Proton NMR Spectra of Monomeric Iron Porphyrins ...................... 32 Proton NMR Resonances for Bis(imidazole)iron(III) Porphyrin Radicals ................................................................................... 33 Resonance Raman Frequencies (cm‘l) for the Parent CoOEP and its Oxidation Products ...................................................................... 67 RR Frequencies (cm'l) of Selected Iron OEP Species ........................ 86 Vibrational Assignments (1450-1700 cm‘l) for Ferric Porphyrins and their n-Cation Radicals ............................................... 86 Resonance Raman Frequencies (cm‘l) and Optical Absorption Maxima (nm) for Parent MOEP and MMeOEC and their Corresponding n-Cation Radicals ...................................... 101 Estimated K(cm'1/A) and A(A) Values for the u and t) Vibrations of the Parent MOEP, MMeOEC and their It-Cation Radicals ................................................................................... 102 Absorption Spectra (nm) of Dicationic Salt of HZTFPP in CHZCIZ ....................................................................................................... 116 Comparison of HzTFPP Absorption Maxima (nm) in CH2C12 with Those of HzTPP and HZOEP ........................................ 117 vii 5-7 6-1 Page Solvent-induced Absorption Maxima Shifts (nm) of TFPP Adducts ..................................................................................................... 118 Changes in the Optical Absorption (nm) in CHzClz of the AgII and CoII Complexes of TFPP Upon One-Electron Metal-Centered Oxidation ..................................................................... 119 Absorption Maxima (nm) of TFPP and TPP Adducts in CH2C12 ....................................................................................................... 129 Absorption maxima and FAB-MS Data on AgTFPP Adducts ..................................................................................................... 130 Absorption Maxima (nm), FAB-MS m/z and 1H-NMR B- proton Multiplicity Number (MN) of ZnTFPP Adducts ................ 134 Stretching Frequencies (cm'l) of (Fe-Cl) and (FeOFe) Vibrations in Several Different Ferric Porphyrin (P) Complexes ................................................................................................ 141 viii LIST OF FIGURES Page 1-1a Oxidation of hemoproteins porphyrin-iron centers in biological systems. ............................................................... 2 1-1b Oxidation mode of photosynthetic chlorophyll a ............................. 2 1-2 Chemical structures of: (a) protoporphyrin IX and its synthetic analogues; (b) meso-tetraphenyl- porphyrin and (c) octaethylporphyrin. ............................................... 5 1-3 The atomic orbital (AO) structure of two HOMOs of porphine. The A0 coefficients are proportional to the size of the circles; solid lines indicate positive values, dashed lines negative. The view is from the positive 2 axis. The straight dashed lines indicate the nodes of the am (It) orbital. ................................................................... 8 1-4 Shape and nodal characteristics of the four essential porphyrin n-molecular orbitals ("Frontier" orbitals; am and am are filled, eg are unoccupied). ...................................... 10 1-5 Typical n-electron energy level scheme of a closed- shell metalloporphyrin: (A) one-electgron energies of highest occupied and lowest unoccupied 1:- molecular orbitals; (B) configuration energies for the lowest n-excitations before (left) and after (right) configuration interaction. (From Wang gt a)” I. Am. Chem. Soc, 1984, m6, 4235.) ................................................... 20 1-6 Visible absorption spectrum of octaethylporphyrinato- nickel(II), Ni(OEP), in CHzClz. (From Kitagawa gt a1_., Struct. Bonding (Berlin), 1987, _6_4, 75.) .............................................. 20 ix 1-8 1-9 1-10 2-3 Page Oxidation of MgOEP in CH2C12. The solid line is the absorption spectrum of Mg(II)OEP, the broken line that of [Mg(II)OEP]+°ClO4'. (From Ref. 16a.) ........................................... 22 Oxidation of MgTPP in CHzClz. The solid line is the absorption spectrum of Mg(II)TPP, the broken line that of [Mg(II)TPP]+°ClO4'. (From Ref. 16a.) ........................................... 22 Upper spectra are those of [Co(III)OEP]7-+°2CIO4‘ (solid line) and [Co(III)OEP]2+°ZBr'(dotted line). The lower spectra are those of the primary complexes of horseradish peroxidase (solid line) and catalase (dotted line). (From Ref. 16a.) ........................................................................... 24 RR spectra of (a) CuOEP; (b) CuOEP+'ClO4'. CHzClz bands are marked with an *. CW laser power 20-35 mW. ........... 36 Visible absorption spectra of CoOEP (Soret = 391 nm) and Com(MeOH)zOEPBr‘ (Soret = 411 nm) in dichloro- methane. .................................................................................................. 49 Electronic absorption spectra of CoHOEP derivatives: (a) CoIIOEP(---); (b) CoHOEP+-C104'(-—-); (c) ComOEP+'2ClO4'(....), in CHzClz; (d) spectral changes on addition of methanol to a methylene chloride solu- tion of (b) producing Com(CH3OH)zOEPClO4‘, total methanol content < 0.01%. .................................................................. 50 Electronic absorption spectra showing the conversion of' CoHOEP+°CIO4' to Com(HzO)ZOEPCIO4' as the tempera- ture is lowered. Solvent, "dry" CHzClz; path length ~ 2mm (EPR tube). ............................................................................................... 52 Electronic absorption spectra of oxidation products of CoOEP. (a) Com(MeOH)zOEPBr'(-); (b) ComOEPBr'(--); (c) ComOEP+'ZBr'(---), the small feature at 401 nm is due to 1% H4OEP2+28r' contamination as discussed in the text. Solvent, dry CHzClz, except for (a) which contains ~ 5% methanol (MeOH). ................................................................................ 54 X 2-6 2-10 2-11 Page 250-MHZ proton NMR spectrum of ComOEPBr' in CDC13 at 25°C. Impurity peaks labeled X. ..................................................... 54 Spectral changes on addition of methanol to a dichloro- methane solution of ComOEP+°2ClO4‘ in the visible region producing Com(CH3OH)ZOEPClO4', total methanol content ~ 2%. ....................................................................... 56 Infrared spectra (in KBr) of CoHOEP and its oxidized products. The porphyrin cation radical diagnostic band is labeled with an asterisk. ......................................................... 59 RR spectra of cobaltous octaethylporphyrin in CHzClz. ............... 61 RR spectra excited at 363.8 nm (~ 35 mW) of CoOEP and its oxidation products. (a) CoOEP; (b) Com(MeOH)zOEPClO4'; (c) CoHOEP+-ClO4’; (d) ComOEP+-2CIO4‘; (e) Com(MeOH)zOEPBr'; (f) ComOEPBr'; (g) ComOEP+°ZBr'. Solvent, dry CH2C12 except (b) and (e) which contain ~ 5% methanol. Solvent bands are marked with an * .................................................................................. 63 Electronic absorption spectra of oxidation products of CoOEP.(a) Com(MeOH)2OEPBr'(-); (b) ComOEPBr' (--); (c) ComOEP+°ZBr' (m), the small feature at 401 nm is due to 1% H40EP2+ZBr’ contamination as discussed in the text. Solvent, dry CH2C12, except for (a) which contains ~5% methanol (MeOH). ................................................................................ 64 RR spectra of cobaltic octaethylporphyrin 1t cation radicals in CHzClz. ............................................................................................... 66 Absorption spectrum of FeIHOEP+-(Cl')(SbC16') in CH2C12 ...................................................................................................... 77 Near-UV RR spectra of OEP+°Fem(X)(X') complexes. The laser power was 35-40 mW. The solvent (CHzClz) band at 1423 cm"1 is labeled with an asterisk (*) ............................................ 79 xi 3-3 3-5 4-1 4-2 5-1 Page RR spectra of ferric chloride porphyrins in CHzClz ....................... 81 RR spectra of ferric chloride porphyrin 1t cation radicals in CH2C12- ............................................................................................... 84 RR spectra of model compounds in CHzClz. Conditions: 25-30 mW incident on the sample in a spinning cell at 25°C. .......................................................................................................... 85 Chemical structures of copper(II) derivatives: (a) trans- octaethylchlroin (t-OEC); (b) methyloctaethylchlorin (MeOEC). .................................................................................................. 91 Optical absorption spectra of CuMeOEC(-) and its one- electron oxidation product, CuMeOEC+°CIO4' (—-) in CH2C12 ...................................................................................................... 93 N ear-uv RR spectra of t-CuOEC (a), Cu (b), Co (c), Ni (d) complexes of MeOEC. The laser power was 20-30 mw. The solvent (CHzClz) band at 1423 cm'1 is labeled with an asterisk (*). ............................................................................................... 95 RR spectra of (a) t-CuOEC+', (b) CuMeOEC+°, (c) CoMeOEC+°, (d) NiMeOEC+°. The laser power was 20-30 mw. The solvent (CHzClz) band at 1423 cm"1 is labeled with an asterisk (t). ............................................................................................... 96 Chemical Structures. ........................................................................... 106 250-MHZ proton N MR spectra of t-OEC and its deuterium labeled derivatives in CDC13. S indicates residual undeuteriated solvent and X represents other impurities. .................................................................................. 108 The electron impact mass spectra (El-MS) of trans-octaethyl- chlorin and its deutero derivatives at 70 ev. ....................................... Absorption spectra of tetrakis(perfluorophenyl) porphyrin complexes taken in CHZClz at room temperature: (A) free base; (B) Zn complex, (C) Cu complex; (D) Pd complex. (From Ref. 7.) ........................................ 109 xii 5-3 5-4 55a 5-5b 5-6a 5-6b Page Solvent-dependent absorption spectra of zinc tetrakis (perfluoro-phenyl)porphyrin in visible region: (A) dry toluene; (B) toluene shaken with HZO;(C) 75% toluene and 25% CH3CN; (D) 75% toluene and 25% ethanol; (E) 75% toluene and 25% pyridine; (F) 75% toluene and 25% triethylamine. (From Ref. 7.) ...................................................... 123 Electronic absorption spectra of CuTFPP(--) and Cu(2-B) in CH2C12 ............................................................................................... 124 Absorption spectrum of H Ag(4-B) in CHzClz. .................................... Expanded B-pyrrolic region of 1H-NMR spectrum of Zn(2-B), fraction 2, in CDC13 .............................................................. 127 Expanded B-pyrrolic region of 1H-NMR spectrum of Zn(2-B), fraction 4, in CDC13 .............................................................. 128 Electron absorption spectrum of H2TFPP/ AgN O3 / HOAC reaction products; t<1 h. ”The 593 nm band is due to the B-pyrrolic adducts. "Gouterman reports AgTFPP Am Ex at 416, 536 and 570 nm in ¢CH3 ............................................ 133 Partial FAB mass spectrum of AgTFPP prepared by AgNO3/HOAC method in < 1 hour. The m/z at 1081 (M+,AgTPP) is not shown .................................................................. 134 UV-vis spectra of CoOEPX in CHzClz. ............................................ 143 Absorption spectrum of ComOEP+-2X' (where X = FeCl4' or C1.) in CHzClz. .................................................................... 145 UV-vis spectra of FeOEPX in CH2C12 .............................................. 146 Absorption spectrum of FemOEP+°(Cl')(SbCl6‘) in CHzClz .................................................................................................... 147 RR spectra of CoOEPX in CHzClz excited at 363.8 nm. The solvent bands at 1422, 703 and 283 cm'1 are labeled with an asterisk. ................................................................................... 148 xiii 6-6 6-8 Page RR spectra of FeOEPX in CHzClz excited at 363.8 nmm .............. 150 (a) R spectra of S4/56FeOEPF in CHzClz excited at 363.8 nm; b) R spectra of 54/ 56FeOEPCl in fH excited at 413.1 nm. ................................................................................................ 152 RR spectra of FeOEPX in benzene excited at 406.7 nm. The solvent bands are marked with an asterisk. .......................... 153 RR spectra of ComOEP+°2X' in CHzClz excited at 363.8 nm: (a) X' = FeCl4' ~ C1-; (b) X' = Br‘. .................................... 155 xiv CHAPTER I INTRODUCTION The diversified functions of hemoproteins in biological systems include roles such as the transport of electrons (e.g., cytochrome b5), the transport of oxygen (e.g., hemoglobin) and the catalysis of redox reactions (e.g., horseradish peroxidase and cytochrome P—450).1'3 These proteins contain an iron (III) protoporphyrin-IX as their prosthetic group. Other related systems, containing a saturated porphyrin ring macrocycle, mediate the rapid flow of electrons in photosynthetic reaction centers (e.g., chlorophyll a, a magnesium chlorin) and catalyze the reduction of nitrite and sulfite (e.g., siroheme, an iron isobacteriochlorin).4r'5 Intuitively, it may be recognized that the widespread use of heme family prosthetic group arises from the presence of a macrocyclic unit, which contains an extensive delocalized n-system (e.g., 18 pi electrons) as well as an iron atom coordinated in the central cavity and, that both the iron and the rt-system are capable of rich electron-transfer chemistry and reversible redox reactions, Figure 1.6 The system is tuned for a particular function by means of specific macrocycle peripheral substituent(s), metal axial ligand(s), and the structural influence of the surrounding protein pocket which imposes both conformational and axial coordination restrictions on the encapsulated heme center. For example, horseradish peroxidase (HRP) and catalase (CAT) are two distinct enzymes that react with hydrogen 9/0 HEMOGLOBIN ,. o, z, mocLoem Ft ,_..—-* Fe crrocnnoues a.” _. h:- .- I HZO? "2° 0 "2°: 0: CATALASE a? >_ Z 2; .p" CAT ‘.3. pEROXlOASE HRP 2 RH 2 R' RH 90H "0:5 C02” r r HE ME CY TOCHnOME P «so Fe c. 02' 2... 2H. "20 crrocaaoue Z“? 2“;- Oz ,4¢-,4H‘ ZHZO Figure 1a. Oxidation of hemoproteins porphyrin-iron centers in biological systems. Figure 1b. Oxidation mode of photosynthetic chlorophyll a. peroxide to produce transient oxidized intermediates referred to as HRP I and CAT 1, respectively, which possess two additional oxidizing equivalents relative to the resting ferric state.7 Compound I in both species has been assigned to a ferryl porphyrin rt-cation radical in which one oxidizing equivalent is stored in the form of a tetravalent iron stabilized by oxygen (FeIV=O) and the second equivalent resides in a porphyrin it-cation. In the HRP case, the fifth proximal ligand to iron is a neutral histidyl imidazole, whereas the iron atom in CAT bears a tyrosine phenolate. This distinction in the nature of the heme-iron axial ligand is presumably the principal factor in the functional differentiation of these hemoproteins.8 The macrocyclic 1t- cation radical species has also been observed in the chlorophyll photochemical reaction for the primary products generated by the photooxidation of the chlorins in the photosystem I and II and, possibly in the cytochrome P-450 reaction cycle.9 In order to account for the contribution that each structural parameter makes to the overall reactivity, synthetic analogs of hemoproteins are invaluable in this regard because they provide the chemist with an opportunity to introduce a composite of several different structural parameters in a stepwise fashion and, consequently, clarify the details of biological redox mechanisms that involve hemoproteins iron-porphyrin centers. Additionally, the protein-free heme model complexes are more frequently responsive to detailed scrutiny than the hemoproteins themselves. While use of specific peripheral substitution pattern of protoporphyrin is essential for the elucidation of biological function, its chemical reactivity and asymmetric distribution make this porphyrin a poor substrate for in vitro model studies. Instead, we utilize two synthetic porphyrins, octaethylporphyrin (OEP) and meso-tetraphenylporphyrin (TPP), which, owing to their high solubility, chemical stability and relative ease of preparation, have proved to be exceptionally useful for model studies. The OEP complexes, however, will be the main focus of this study owing to their biologically-relevant substitution pattern, analogous to that of protoporphyrin species, Figure 2. Focusing on these structurally-related porphyrins will demonstrate systematically how the model compound reactivity may be tuned to obtain a desired potential (E0), a desired number(s) of electrons transferred in the reaction sequence,or a specific ultimate product (e.g., change in the valance of the central metal or generation of a n-cation radical). Research Objectives This study aims to utilize synthetic porphyrins and their derivatives to establish vibrational, electronic and structural characteristics of the oxidized states of hemoproteins iron-porphyrin centers. The emphasis will primarily be placed on the porphyrin-based n-cation radical state which serves as one of the two oxidizing equivalents in the reaction cycle of peroxidases, catalases and photosynthetic chlorophylls. At the time that the present work began, the redox behavior of metalloporphyrins had been widely investigated and there were already numerous reports available on the visible absorption, ESR, magnetic susceptibility and NMR properties of these radicals;10‘13 however, their vibrational characteristics had not been studied in any detail. This may partially result from the limited stability of these radicals, their preparation under conditions (e.g., solvent, co-solvent and concentration) COZH COZH (a) (b) (c) Figure 2. Chemical structures of: (a) protoporphyrin D( and its synthetic analogues; (b) meso-tetraphenyiporphyrin and (c) octaethylporphyrin. that may not be suitable for vibrational studies, and the time required to scan a spectrum when a conventional spectrometer is used. Resonance Raman (RR) spectroscopy has been utilized as a powerful probe, in the last few decades, to study hemoproteins structures, functions, and mechanisms. A major advantage of R is its selectivity for individual chromophoric species in a multicomponent hemoprotein system. Since the interpretation of RR vibrational frequencies relies heavily upon previous studies of model compounds that were designed to mimic the structure of molecular species thought to exist in the more complicated protein sample, this prompted us to investigate vibrational characteristics of synthetic metalloporphyrin n-cation radicals. A systematic study of these radicals will demonstrate the sensitivity of RR frequencies to the vibrational, electronic and structural changes accompanied by removal of an electron from the porphyrin-based it-system. This will, in turn, aid detection and assignment of this particular mode of oxidation in the catalytic cycles of hemoproteins currently under investigation. The RR data obtained in this study are intimately supported by other spectroscopic methods, in particular, optical absorption, ESR, magnetic susceptibility and IR data. It is the objective of this chapter to present a general background of the structural parameters involved and the methods commonly used in the study of metalloporphyrin redox reactions. Porphyrin n-Cation Radicals The striking feature of the porphyrin ring is its ability to undergo facile oxidation under chemical or electrochemical conditions to form a 1t—cation radical. Probably the first direct evidence for the existence of porphyrin 1t- cation radicals was Commoner's observation of a rapidly decaying, sharp ESR signal at g = 2.00 in photosynthetic material.” Another early report on signals that were similar but of high stability was made for solutions of TPP.15 Easy access to stable cation radical as well as their detailed analysis, was achieved when ESR spectra, absorption spectra and redox potentials were quantitatively related to the oxidative formation of these radicals and crystalline material was obtained.16 Gouterman17 has shown that in a metalloporphyrin (MP), the highest filled orbitals, am and am, are nearly degenerate, Figure 3. Removal of an electron from these orbital produces a n-cation radical and the remaining unpaired electron occupies either an am or an orbital with zAlu or 2A2u ground state (D411 symmetry), respectively. Based on recent calculation by Zerner,18 the two ground states differ in energy by about 4 kcal/ mole; thus, small perturbation such as changes in peripheral groups, axial ligand(s) or the central metal could determine the symmetry of the ground state. Moreover, MO calculation as well as ESR data showed that in the 2A211 state, high spin density appears at both the meso-carbon and nitrogen atoms, while in the 2A1u state, spin density is primarily localized on the a carbons, Table 1.19 For example, the ESR evidence confirms that, indeed, the n-cation radical of MgOEP occupies a 2A1u and that of MgTPP a 2A2u state.20 As a rule, the am orbital is found to be higher than the a1u orbital. Taking into account the fact that am orbital has its maximum electron density on the nitrogen and meso-carbon atoms, we can assume that addition of meso-substituents (e.g., phenyl groups in TPP) will lead to an increase in the Hm-l ”32“ int" Figure 3. The atomic orbital (A0) structure of two HOMOs of porphine. The A0 coefficients are proportional to the size of the circles; solid lines indicate positive values, dashed lines negative. The view is from the positive 2 axis. The straight dashed lines indicate the nodes of the am (it) orbital. Table 1. Spin Odie! in Porphyrin Cation Radicals Morn (calculated) (calculated) C-l - 0.0094 0.098 l C -2 0.0l 34 0.0262 C-5 0. l932 0.00l 2 N 0.049 0.000 (From Ref. 19.) energy of this orbital to an extent that it becomes higher than the energy of the am orbital. In this case, the electron abstraction will take place from this orbital to form a 2A2u n-cation radical. Indeed, almost all metallotetra- phenylporphyrin cation radical display a 2A2u ground state. In the OEP case, the introduction of alkyl substituents in the pyrrole rings probably results in the alu orbital being higher than the am level. The electron density will therefore be removed from the orbital with am symmetry to form a zAlu cation radical. However, several investigations have demonstrated that the OEP-type system is more responsive to the change in the ground-state symmetry as a result of perturbation caused by metal or axial ligand, than the corresponding TPP complexes.21 Metal-Porphyrin Interactions The role of the centrally coordinated iron in hemoprotein catalysis is to hold the reacting species in conformations favorable for reaction and act as a source of, or sink for, electrons. Calculations on model compounds have shown that metal interacts with the porphyrin ring in two ways, which are referred to as inductive and conjugative effects.”- 1. Inductive Effect: This mode of interaction results from the change of electron density at the nitrogen atoms as the result of substituting different metals in the central cavity. This perturbation affects the energies of the doubly degenerate LUMO eg(1t*) and HOMO a2u(1t) but not alum) orbital because the latter has nodes through the central nitrogen atoms, Figure 4. As the central metal becomes less electronegative, the sigma electron shell shifts from the metal towards the ring which raises the eg and am orbitals by inductive effect. 10 Figure 4. Shape and nodal characteristics of the four essential porphyrin 1:- molecular orbitals ("Frontier” orbitals; a2“ and am are filled, e8 are unoccupied). 11 2. Conjugative Effect: This effect results from the interaction of metal P,t orbitals with those of porphyrin it-electron system (e.g., eg and am). For example, as the central metal becomes less electronegative, that part of the pi electron density located in the metal 4P2 orbital can move back into the ring causing an increase in the energy of the an orbital at the expense of a decrease in conjugation between metal and porphyrin pi orbitals. Site-Selective Oxidation: Metal- vs. Porphyrin-centered Oxidation Induction Metalloporphyrin redox reactions may be classified into three categories: 1. Reversible changes in the formal oxidation number of the metal ("inorganic redox reactions"), 2. Reversible changes of the oxidation state of the porphyrin ("organic reversible redox reactions") and 3. Both metal and the porphyrin ligands undergo reversible redox reactions. As was previously noted, several structural parameters contribute to the selective oxidation reactions of the metal and porphyrin centers. These are: 1. the nature of the central metal, 2. macrocycle peripheral substituent(s), 3. metal axial ligand(s), 4. solvent (or medium) and 5. temperature. The role of these parameters will be discussed individually in the following section and examples will be given. 12 Effect of the Central Metal: The redox chemistry of metalloporphyrins is dominated by strong o-electron donation from the nitrogens to the metal ion. Metals of low electronegativity and low oxidation number activate the porphyrin ring with respect to oxidation and deactivate it with respect to reduction. The opposite is true for metals of high electronegativity and/ or oxidation number. Such reversible reactions of the macrocyclic ring are of importance in biological systems containing magnesium metal; namely, photosynthetic chlorophyll a. On the other hand, electron-rich transition metal ions (e.g., iron) with formally high oxidation numbers are stable in the porphyrin cavity and tend to transfer some electrons to the axial ligand. This effect is of biological importance in hemoproteins, which function in electron transfers and oxygen activation. Porphyrin Peripheral Substituent(s): The one-electron oxidation products of the MP's are highly reactive species having redox potentials between those of 12 and Brz (~ 0.5 - 1.2 volts). Experimentally, it is found that oxidative titration in organic solvents with a stoichiometric amount of oxidant are only possible for complexes with midpoint potential below 0.3 volts (e.g., magnesium octaethylchlorin, MgOEC; ZnOEC and BChl a) and iodine as an oxidant. In all other cases, an excess of oxidant is needed to complete the oxidation of MP probably because the organic solvents, or reducing impurities contained in them consume oxidant above 0.5 volts. A range of oxidants may be used to produce n—cation radicals including 12, Brz, CuBrz, NBS, DDQ, Fe(ClO4)3, FeC13, pheno+- SbCl6', ¢3N+X’, CeIV salt, XeF223. In special cases, stable cation radicals 13 of porphyrin itself may be used to induce oxidation in another porphyrin system such as ZnTPP+-ClO4'(81 /2 = 0.77 volts vs. SCE) and (FeTPP)zO+°ClO4'(81/2 = 0.84 volts vs. SCE).24 Recent findings concerning the role of the propionate and vinyl groups of iron (III) protoporphyrin-IX in cytochrome b5 highlights the significance of macrocycle peripheral substitution pattern. This study, reported by Reid e_t a_l,25 indicated that heme proteins modulate the reduction potentials of their heme redox centers by constraining the position of the vinyl groups with respect to the plane of the heme and thereby affecting the electron-withdrawing ability of the vinyls. By forcing the vinyl substituents into the plane of the heme prosthetic group, their electron-withdrawing ability is enhanced, and the reduction potential of the heme center is increased. In another study, they also suggested distinctly different roles for heme propionate groups, one partially responsible for determining the reduction potential of the iron center, and the other group is involved in recognition of and interaction with heme proteins that are physiological redox partners of cytochrome b5. As demonstrated by model studies, in general, the presence of electron-donating or electron-withdrawing substituents modify the electron density at the central metal and lead to shifts in redox potentials, the magnitude of which are dependent on the substituents and their orientation with respect to the macrocycle ring plane. For example, metallochlorins are oxidized by as much as 0.3 volts lower than metalloporphyrins, and bacteriochlorins are again 0.3 volts lower. This trend can be rationalized by noting that in the B-hydrogenated l4 porphyrins, one or two of the electron-withdrawing pyrrole units, which can be formulated as pyrrole anions (pk~16), are replaced by pyrroline units, which function as aza bridges (pk~2) in the inner conjugation path. Therefore the ligand becomes less basic and the oxidation potential of the central metal increases; therefore, it becomes easier to remove an electron from the porphyrin ligand itself.26 Metal Axial Ligand(s): Metalloporphyrins can bind one or two ligands at the axial coordination sites and these ligands can potentially dictate the chemistry of the heme group. The heme in the electron carrier hemoproteins is exceptional in that it has two strong axial ligands and generally does not bind molecular oxygen or peroxides. The heme in all other hemoproteins has one accessible coordination site which allows it to react with peroxides and other ligands (Table 2). Table 2. The Nature of Axial Ligand(s) in Some Hemoproteins l ‘ Hemoprotein Axial Ligand(s) Cytochrome P-450 Cystein thiolate (1) HRP Histidine Imidazole (1) CAT Tyrosine Phenolate (1) Cytochrome c peroxidase Histidine Imidazole (1) Cytochrome b5 Histidine Imidazole (2) Cytochrome c Histidine Imidazole (1) Methionine Methylthioether(1) The number in parenthesis indicates the number of axial ligand(s) coordinated to the heme-iron center. 15 The nature of the axial ligand can trigger metal vs. porphyrin-based oxidation, influence the ground state symmetry of the porphyrin radical (2A2u or 2Am), alter the spin state of the central metal and disturb the conformation of the macrocycle. For example, the site of oxidation of ruthenium(II) porphyrins (metal or porphyrin) is dependent on the nature of the axial ligand. When pyridine occupies the extraplanar sites, oxidation occurs at the ruthenium ion. If CO is one of the extraplanar ligand, the site of oxidation is at the porphyrin ring. This change has been attributed to stronger ruthenium back-bonding to CO compared to pyridine, which stabilizes metal d1; orbitals relative to the porphyrin 11: levels; the latter then becomes the valance MOs of the system.27a It has also been shown that in iron(III) tetramesitylporphyrin, TMPFeIH(L), strongly basic ligands such as L = o‘,OH',OCH3',F' and = O favor the formation of iron (IV) species, whereas weakly basic anions such as L = ClO4‘, Cl' and imidazole cause porphyrin cation radical formation.27b As discussed earlier, the species COIHOEP+°2X‘ can be made to occupy either of the two ground states (2A1u vs 2A2“) as a function of its axial ligands. When the dibrornide (2A1u state) is treated with AgClO4, the bromide ligands are removed and the resultant diperchlorate salt now shows characteristics of the 2A2u state.16a The variation of the axial ligand can result in a change in Aspin state, particularly of iron porphyrins, and concomitant structural changes.28 High-spin hemes (S = 5/ 2) are five-coordinated and exist in 16 square-pyramidal conformation. The iron atom is displaced from the porphyrin plane toward the axial ligand (e.g., X‘ = Cl',N3') which makes the binding of further ligands unlikely. On the other hand, the formation of two strong axial bonds (e.g., bis imidazole derivative) involves an in-plane position of iron and a low-spin state (S = 1/2). Thus, the obligatory structural changes that accompany the low-spin to high-spin transition is an expansion of the porphyrin core size (e.g., PFem(Im)2+ —> PFem(DMSO)2+). High-spin Fe(II) and Fe(IH) porphyrins are both too large to fit into the central cavity of the porphyrin ring and they are forced out of the porphyrin ring by ~ 0.4 and 0.5 A, respectively. These out-of-plane structures exhibit significant doming and ruffling of the porphyrin ring.29 Solvent (or Medium): Protic solvents (e.g., CH3OH) slowly destroy the cation radical and may only be used as co-solvents in oxidation reactions, when the electron transfer is very rapid. Although, the stabilizing effect of nucleophilic solvents on the oxidation of certain metalloporphyrins (e.g., MgOEP) favors the use of solvent mixture such as CHC13- CH30H(4:1);10 however, the electrophilic nature of the porphyrin cation radicals may cause nucleophilic addition of the solvent to the bridged meso carbon of the cation radical and the subsequent formation of an isoporphyrin complex.30 Thus, care must be taken to avoid occurrence of this side reaction. Solvent effects on the porphyrin redox potentials are small. However, solvent effects on metal redox potentials can be extraordinarily 17 large. For example, the CoIH/CoII redox couple involves low-spin d6/ d7 configurations. Strongly binding axial ligands (solvent molecules) destabilize the dzz electron (in d7) and favor oxidation to low-spin d6 species. Thus, in this case, the potential shifts negatively with increasing donor strength of the solvent (e.g. DMA = DMF< DMSO < pyridine).31 Solvent may also play a role in the selective formation of metal- or porphyrin- centered oxidation product. The first oxidation of CoOEP in dry CHzClz, for example, involves formation of a Co(II) porphyrin cation radical while the second and third oxidations involve abstraction of electrons first from the Co(II) center and then from the porphyrin 1t- system. This ordering in the sites of oxidation is reversed in coordinating solvents or in the presence of nucleophiles. In the latter solutions, the first oxidation of CoOEP invariably involves the Co(II)/Co(III) transition.32 Temperature. Oxidation of NiTPP at room temperature results in the abstraction of an electron from a porphyrin-centered 1t orbital to form a n-cation radical. This species, NiTPP+; displays an exceptional thermal behavior. As the temperature is lowered, an internal electron transfer occurs, with the hole moving from the periphery to the metal to generate a N i(III) species.3 Although the structure of protein determines the heme-iron spin equilibrium in hemoproteins through control of axial 1igation,34 a thermal spin-state equilibrium can be observed in a model heme system whose axial ligation is characteristic of these proteins.35arb Several studies have indicated that the nature of the axial ligand is an important 18 factor in determining the high-spin or low-spin character of the iron center. Weak field ligands such as F' form high-spin, six-coordinate iron (III) complexes; strong field ligand such as CN’ and imidazole form low- spin hexacoordinate iron (III) complexes, and ligands such as N 3‘ having intermediate field strength form complexes that exhibit temperature- dependent, spin-state equilibrium.36 The position of the equilibrium at a given temperature is sensitive to the immediate environment of the iron(III) porphyrin cation. For example, OEPFeHI(py)2+ CIO4" show a magnetic moment at 77°I< corresponding to an almost pure low-spin state (S = 1 / 2) and a moment at room temperature appropriate for 1:1 mixture of the low-spin and high-spin (S = 5/ 2) states.3513 Differentiation Between Metal vs. Porphyrin-based Oxidation Reactions - Spectroscopic Methods 1. Optical Absorption Spectroscopy: The oxidation reactions of metallopor- phyrins usually can be followed visually: a neutral metalloporphyrin gives a reddish color solution; oxidation most often changes this color to dark brown or green if the reaction is porphyrin-centered, whereas the red color maintains if the reaction proceeds at the central metal“), 32 Since MPs and their oxidized products display characteristic electronic absorption spectra, the following section will briefly address the optical properties of these species. A typical absorption spectrum of the metal complex of the biological-type (spyrrole-substituted) porphyrin (e.g., OEP) features two visible bands (at or 00,0 and B or Q03 bands) of moderate intensity (8 z 19 102-103 M'lCm‘l) in the 500 - 600 nm region and a very strong (8 z 104- 105 M'lCm'l) band, called the Soret (also 8 or B) band near 400 nm, Figure 6. The visible and near-uv spectra of MPs can be interpreted within the context of the four-orbital model proposed by Gouterman and coworkers.37 The electronic transitions that give rise to the characteristic spectra of MPs arise by excitation from the two highest filled orbitals (HOMO's), the nearly-degenerate a1u(1t)and a2u(1t) levels, into the doubly-degenerate lowest unfilled orbitals (LUMO's), eg(1t*) levels, (D4h symmetry). For the porphyrin ring, the lowest singlet excited configurations, 1(a2u,eg) and 1(a1u,eg), are of the same symmetry (Eu) and of similar energy, Figure 5. These transitions have strong electronic interaction between them and are mixed by configuration interaction to yield the relatively weak visible Q0,0 band in which the transition dipoles nearly cancel, and the intense near-uv Soret band, in which the transition dipoles of the two configurations add. The closer the degeneracy of the configurations '(a2u,eg) and '(alu, eg), the weaker is the Q0,0° In some instances (e.g., CoTPP), the Q0,0 band essentially disappears. In addition, the Q0,l vibronic overtone is also active and appears as an additional peak on the high-energy side of 00,0 transition. The Q0,1 band arises from vibronic mixing of the Q0,0 state with the B state and is constant among porphyrins.38 The assignment of this band is based on the constancy of its energy separation from the Q0,0 band and its nearly constant intensity. Comparison of the optical spectra of a number of n-cation radicals16a reveals that they fall in two categories typified by either 20 A 3 Po“ Alter Orbitoi Energies Configurations C! a ”"1:— I (o.‘. s: i—‘ . .. I T“ cum I I Ground Slots ‘.to E Figure 5. Typical zit-electron energy level scheme of a closed-shell metalloporphyrin: (A) one-electgron energies of highest occupied and lowest unoccupied n-molecular orbitals; (B) configuration energies for the lowest n-excitations before (left) and after (right) configuration interaction. (From Wang gt 31., I. Am. Chem. Soc, 1984, 1% 4235.) Ni(OEP) ABSORPTION Xl/2 l L I l L l 400 500 600 WAVELENGTH (nm) Figure 6. Visible absorption spectrum of octaethylporphyrinatonickeKII). Ni(OEP), in CHZClz. (From Kitagawa gt 3], Struct. Bonding (Berlin), 1987, fi, 75.) 21 MgOEP, Figure 7, and MgTPP it-cation radicals, Figure 8. The MgOEP radical has a visible spectrum that is characterized by a major absorption peak near 700 nm with a high-energy shoulder, while the MgTPP radical exhibits a broad, near featureless absorption in the region 500-700 nm. These differences have been ascribed to two close-lying ground states of the radicals: 2A1u (class 1) for MgOEP+' and 2A2“ (class 2) for MgTPP+° The preceding interpretation has been utilized for some time as the optical criterion by which to distinguish between the two possible ground states. Qualitatively perhaps, it is still useful, however exceptions have been found and caution must be used in making any definite assignments that are based on the optical characteristics. For example, Godziela gt a_l. recently showed that CuOEP+°, previously taken as an example of 2A2u ground state,16a exhibits NMR chemical shifts characteristic of a 2A1u radical.39 A recent study by Zerner and coworkers has provided a firm foundation for a quantitative interpretation of MP n-cation radicals.18 This work demonstrates that 2A1“ cations lie lower in energy than the 2A2u cations by only 4 kcal/mole, consistent with the findings that both radicals are found depending upon substituents and solvent. The visible region of the spectrum of the 2A1u species is predicted to consist of three separate bands decreasing in intensity with increasing energy, while that of the 2A2u species is calculated to consist of three allowed transitions of nearly equal intensity, in accord with the experimental findings. The Soret region of the 2Am ions is dominated by at least four allowed transitions spread over 100 nm, while that of the 2A2u radicals contains at least 22 gore-0’ '0 0-. coco-o..-------- 0" a a Figure 7. Oxidation of MgOEP in 01202. The solid line is the absorption spectrum of Mg(II)OEP, the broken line that of [Mg(mosrlraoat (From Ref. 16a.) me (out Figure 8. Oxidation of MgTPP in CHzClz. The solid line is the absorption spectrum of Mg(II)TPP, the broken line that of [Mg(II)TPPP'ClOf. (From Ref. 16a.) 23 three allowed bands, but the intensity in this region is dominated by only one of them. Thus, an absorption spectrum with two sharp bands in the visible region and a narrow Soret band indicates a pyrrole-substituted MP with the porphyrin ring in its zero oxidation state. If this spectrum is found in an oxidation product from a MP, then oxidation has occurred at the metal (e.g., AgHOEP -> AngEP).40. If, however, the electronic spec- trum broadens considerably after oxidation, then the porphyrin ligand might have reacted (e.g., CuHOEP —> CuIIOEP‘l").16a As a matter of fact, similarity between the optical spectra of catalase and horseradish peroxidase compound I with those of CoIIIOEP+° 2Br" and ComOEP+° 2ClO4', respectively, led to the notion that these complexes may be characterized as porphyrin it-cation radicals and that, CAT I exists in the 2A1u ground state, whereas HRP I exists in the 2A2u state. It was also speculated that differences in reactivity between the two enzymes are the result of differences in the ground states, which may, in turn, be caused by minor changes in the axial ligation, as demonstrated by the optical characteristics of the two cobalt octaethylporphyrin cation radicals, Figure 9.16a ESR Method. If only one unpaired electron is found in the product, then its location is easily determined by examination of line width, g-values and hyperfine structure in the ESR spectrum. Typically, the ESR experiments yields two possible results: 1. An isotropic signal characteristic of an organic free radical around g = 2.00 and of line width less than 12 G (e.g., ZnTPP‘“-).19 24 ex IO"5 560 600 760 )\,nm Figure 9. Upper spectra are those of, [Co(III)OEP]2*'2ClO4' (solid line) and [Co(III)OEP]2*"28r'(dotted line). The lower spectra are those of the primary complexes of horseradish peroxidase (solid line) and catalase (dotted line). (From Ref. 16a.) 300 4b0 25 2. An anisotropic signal of line width > 50 G with various g values typical of transition metal ions (e.g., CoIIITPP*’°).41 Copper porphyrin cation radicals, on the other hand, are more complex owing to the presence of one odd electron on the central metal (3dx2-y2 orbital) and the other on the porphyrin ring. Therefore, their spin state is either singlet or triplet depending on which is lower in energy. Since no triplet ESR spectrum had been detected for some time for any copper porphyrin radicals, it was speculated as being due to a singlet ground state or to line broadening of triplet spectra.10 However, Konishi and coworkers recently reported the first detection of a triplet spectrum of a copper porphyrin cation radical produced in tetrachlorethane matrix at 77°I< by 8 radiolysisflr2 The ESR spectra obtained for MR“ can be classified into two types, exemplified by 2A2u and 2A1u ground states. The ESR analysis of the 2A2u state reveals: 1. Large spin density at the meso carbons which can be transmitted to the meso substituents (e.g., phenyl group). 2. Enough spin density at pyrrole nitrogens to cause nitrogen hyperfine splitting (1-2G). 3. Zero spin density at B-pyrrole positions. 4. Enough spin density at the metal to cause metal hyperfine splitting (e.g., azn = 1G). 5. Charge density transmission to the counterion (e.g., ZnTPP+-(Py)ClO4‘). In the 2A1u state, on the other hand, small hyperfine splittings from the meso protons are observed (e.g., an = 1.48G for MgOEP+°ClO4'). This 26 ground state also does not display any splitting due to the nitrogens, central metal, or counterion and the electron density is mostly localized on the a carbons. 3. Magnetic Susceptibility Measurement: When both the metal and the porphyrin ring contain unpaired electrons, magnetic susceptibility measurements by using a SQUID susceptometer in the solid state43 or Evan's NMR shift method in solution44 help to locate the site of a reaction. The theory of spin state in these MP4" is based on the occupations and symmetries of the so-called magnetic orbitals, i.e., those orbitals on the metal and the ligand that contain an unpaired electron. MP4": complexes whose half-occupied d orbitals are strictly orthogonal to the porphyrin n-radical orbital are expected to show ferromagnetic coupling (like Hund's rule). Thus, the spin state of higher multiplicity will be lower in energy. For example, in a strictly planar environment of D4h symmetry as well as in the absence of any complication from aggregation, CuTMP+°SbCl6’ features orthogonal copper dx2-y2 and porphyrin am or a2u orbitals and shows an S = 1 (i.e., triplet), ueff = 2.9u3, ground state.45 On the other hand, complexes whose metal and ligand magnetic orbitals are not strictly forbidden by symmetry to overlap are expected to show antiferromagnetic coupling (or bond formation). The spin state of lower multiplicity will be lower in energy. This is exemplified by the CuTPP+' SbCl6' crystal which shows an S = 0 (i.e., singlet) ground state. This crystal exhibits a saddle-shaped conformation in which nearly planar pyrrole rings tilt up and down and the phenyl groups become nearly coplanar.4r6 Although this 27 conformation allows the appearance of tightly associated pairs of cations, the face-to-face structure of the dimer lacks any recognizable specific HOMO-LUMO interaction or the popular but unproven intermolecular metal-nitrogen interaction. The large Cu-Cu separation (5.43 A) also makes it unlikely for the trivial d-d coupling between cofacial dx2-y2 orbitals. For similar reason, the intermolecular d-n: coupling is also considered unlikely. This leaves an intramolecular d-n: coupling within each CuTPP+° molecule as the likely source of complete diamagnetism. Thus, Reed and coworkers‘lr6 suggested that intramolecular metal- porphyrin spin coupling occurs as a result of the lowering of the local symmetry from D4h to Cs which destroys the orthogonality of magnetic orbitals. Under CS symmetry, the dx2-y2 and the am orbital overlap become symmetry allowed and the resultant spin pairing may be viewed as copper porphyrin bond formation with the two electrons spin paired in the bonding molecular orbital. In solution, however, the monomeric CuTPP+° is planar and strict orthogonality of the magnetic orbitals gives rise to the paramagnetic state. A yet alternative mode of coupling is demonstrated by CuOEP+° whose nearly planar conformation is prone to the formation of tightly- spaced dimers (Cu-Cu separation of ~ 4A) at low temperatures or in solid state and at high concentration in solution.45 This can reduce the moment of the monomeric CuOEP+' triplet ground state (ueff = 2.9 113) as a function of the extent of association into dimeric units or higher order aggregates. The driving force for dimerization is probably best 28 viewed as a case of weak 1t—1t bond formation between two half-occupied porphyrin HOMOs. Indeed, N iOEP+° is diamagnetic from presumed 1t-1t coupling.46 Electrochemical Regularities: A systematic voltammetric investigation of the redox potential of metal complexes of octaethylporphyrin including all members of the first transition series has yielded some useful electrochemical regularities that provide additional support for the site of oxidation in metalloporphyrins.47 1. The differences in potential between the first oxidation and reduction of the porphyrin ring are, with two exceptions (Mn, Mo), constant: on_ - Bred. = 2.25 i 0.15V 2. The differences in potential between the removal of the first and second electron are also constant: E20,, - Elox. = 0.29 3: 0.05v These empirical rules together with chemical and spectroscopic data may be used to correlate various redox steps of a metalloporphyrin system with metal vs. porphyrin-centered reactions. The more interesting of these two electrochemical rules is certainly the first one: regardless of any intermediate change in the oxidation states of the central metals, the difference between the highest occupied and lowest unoccupied orbitals is a constant 2.2 eV in the ground state. With stable divalent metals, linear correlations between redox potentials and metal electronega- tivities have been demonstrated.26 The fact that a rise in oxidation potential is accompanied by a lowering of the reduction potential clearly points to the inductive effects of the metals on the porphyrin n—system. 29 Voltammetric waves corresponding to potentials which are far away from these patterns are expected to be associated with reactions of the central metal ions. Other Physicochemical Methods: In addition to the more conventional methods of investigation described above, metalloporphyrin redox reactions have recently been studied by other techniques including NMR, IR, and resonance Raman (RR) spectroscopy. The utility of these methods will be briefly discussed in the following sections; however, RR vibrational analysis of macrocycle-based oxidized hemoprotein models represents the main subject of this dissertation and will be covered in detail in the following chapters. Proton NMR of the Metalloporphyrin it—cation Radicals: The salient features of 2A2u porphyrin it-cation radical are distribution of a substantial amount of positive spin density at the meso carbons and the pyrrole nitrogens, while the radical orbital in zAlu cations has a node at these atoms (D4h symmetry).19 Since the isotopic shift of the meso proton is proportional to the meso carbon spin density according to the McConnell equation,8 the typical 2A2u radical is expected to induce a large upfield contact shift for the meso proton whereas the zAlu radical should exhibit a downfield bias.4*9 In the TPP complexes, the n-cation radical character, described as an an radical type, is strongly implicated by a large upfield and downfield spread of the phenyl proton signal, Table 3. The most striking difference between high-spin and low-spin iron(III) TPP radical is the fact that the signs of phenyl resonances are 30 reversed for the two species. Thus, phenyl ortho— and para-proton signals are far downfield for the high-spin chloroiron(IH) species, Table 4; but, shifts of the same magnitude, but upfield, are seen for the bisirnidazole complex, Table 5. N 0 description is offered for this reversal phenomenon; however, it is suggested that the phenyl shift pattern of the high-spin iron(III) porphyrin radical is the exceptional case and that the pattern for the low-spin iron(III) radical parallels the shift pattern of several other recently identified MP4" (e. g., CuTPP+', Table 3). A shift of the OEPFeCl methine protons upon oxidation, Table 4, is consistent with the expected shift direction of known alu radicals, also demonstrated by CuOEP+' C104“, Table 3. Another interesting observation made by Morishima and co- workers is concerned with the non-Curie law behavior of the temperature dependence of meso-D resonance shift in ComOEP'l'QX' (where X = Br',ClO4“).51 They interpreted this behavior in terms of thermal equilibrium between 2A1u and 2A2u states. If we assume as they suggested, that the am orbital possess slightly higher energy than the an orbital in biological-type porphyrins, the thermal mixing of the 2A2u state increases with increasing temperature to afford more positive spin density at the meso positions, and eventually inducing more upfield isotropic shift. IR of Metalloporphyrin n-cation Radicals. This method provides a diagnostic tool for detecting porphyrin It—cation radical character, as was first reported by Shimomura and Goff.52 For the TPP radical complexes, an intense new band appears in the 1280 cm"1 region. Likewise, oxidized 31 Tab Lhasa-dMNHIMMI¢Ce~sd$~hmsuMOerMMnW m est-st W pet-vi inn-I- M- has <3“:- ti-s 4“. «'1‘ {Hr an)“. as us I?) 1.11 m a: (movie I It I w (m (OEPiAs' 1" ll 1.)! (“l (oer-4.34.! -Is.I (on; i1.) um iflPP-d.)AeiOO.' us (mi "endgame: 9.0 (up 4.)! (m (We‘ 1.“ (lat 1.10mi 1.31 (i0? is IMJ-(OCHMI'MCE u ms) 3.“ my 1.” (i n- x meat x 1.“ (m m ()9) at (Tn-med u unset Image-loo: I 3.4 I i711 l("'-J.)Cilmg' «.10 not us In) -I.oI (m IOflicv x I L: (I213) LII Mi (OH-axe 4.1 (20st ".3 (I!) Itoeugcotoo: «.I (130) In (I21) (ETIOKI' x ”J (tom u (Jon i.“ my (filo-JuiCe’ 4.2 (203) it.) (“i u (m l(ETlO-J,.)CulClO.' «sum it? user In (th awmmmatuvmmmmmmmmmrs. Linewidthminfi parentheses,areglveninHL'X'aitriesuidlatesigmlsnotdetectedduetolargelimwidthand/or overlap with othersigmls. b(231303 solver". cCI-12Clzsolwent. dCD202 solvait. 'p-CH3 sight. ftn- OCi-l3 siyul. 89-00-13 sight. (From Ref. 39.) 32 Table 4 . Inseam! Soecoadltmieirea Porphyr'us‘ canoe pyrrole «the meta para Cit. ammo ”.4 -s i3.J. 6.33 In (TMI’thOOJ “.i 17.6. -Iz.4 29.5 34.4 macaw-ac: 191 -s In. $.2i ii.’ mum-warm.) 59.9 out 4.3 4.1 ti‘iOi‘ amt IIuI‘ (mt rrru-ocngncmoom as 41.3 4.2 «a coated Cit, Cit, mesa (OEPiF-O‘ 0.1. 39.3 s.“ -54 (OEflFeOiOOJ‘ 10.5. 29.6 3. l6 -ll 8Except as noted oxidized species in CD202 solvent, others in CDCl3 solvent. iron porphyrin - 10 mM. 26°C, Me45i reference. bLine widths at half height. scum solvent. dao-c; Itau-c. (From Ref. 49.) Tale S.Protoa Null Reno-ans for listinirhaoieiitofllll) Wrist lladieats‘ TPP- TPP- proton «pocm‘ 17" (Lu-oat r 09‘ ETIO‘ pyrrole ~32.7 «on 40.4 o-oitenyi -36.J -Il.7 walnut 24.8 30.6 26.7 "heart -22.l pOCH, i2.9 9.5 , I18 tin. CH, SIJ $0.4 tine CH, till W2 solvent. 0.01 M iron porphyrin, 0.5 equivalent of imidazole. referenced to (CH3)4Si; downfield shifts are given positive sign. 968°C. C-3O‘C, 1.0 equivalent imidazole present.d-51'C; ethyl-CH3 groups were obscured by the solvent signal; the mesa proton signal was not detected. (From Ref. 50.) 33 OEP radicals display a diagnostic band in the 1550 cm'1 region. Other workers53 have criticized the use of this band as a diagnostic criterion, and produce evidence suggesting that it could arise from the trapped chlorocarbon solvent molecules in the crystalline cation radicals. Hinman and coworkers have recently studied the cation radicals of several metallotetraphenylporphyrins and concluded that the formation of the n—cation radical results in two characteristic absorbance decreases near 1485 and 1600 cm'l, and five characteristic absorbance increases near 1005, 1225, 1285, 1350, and 1415 crn-1.54 Investigation of the cation radicals of OEP complexes was also recently carried out by Itoh and coworkers.55 The tentative assignments of infrared bands of these radicals indicate frequency shift patterns relative to the unoxidized molecules that can be used for determining the ground electronic states of the n-cation radicals. Resonance Raman. The application of RR spectroscopy to biological systems is an area of intense activity. RR spectroscopy is potentially able to furnish information about a molecule and its environment under experimental conditions which are not favorable to classical vibrational spectroscopy (low concentration, complex media, aqueous solution). Due to the phenomenon of resonance enhancement, specific parts of complicated molecules (e.g., hemoproteins) may be probed without interference from other parts of the molecule or from the solvent. The utility of this technique as a structural probe of the chromophore in hemoproteins has stimulated studies on protein-free metalloporphyrins, yielding a great deal of information about structural, spectroscopic and physical properties.563 Laser excitation into porphyrin Soret or visible bands enhances the scattered intensity by 103 - 104 and provide selective excitation of porphyrin vibrational modes. Empirical correlations between the positions of Raman lines and metal oxidation, ligation and spin state and the core size (center-to-nitrogen) distances exemplify some of the information accessible by RR spectroscopy.56b Identification of peripheral substituents and the dependence of the depolarization ratio II p = m of certain lines as a function of the wavelength of incident radiation is interpreted as evidence for electronic symmetry reduction, i.e., CuOEP(D4h) —) CuOEC(C2v).57 The dominant RR bands in the 1000 - 1700 cm'1 region correspond to the in-plane stretching of C-C, C-N partial double bonds and the bending of C-H bonds. At low frequencies, < 1000 cm'l, the RR spectra are dominated by out-of-plane modes of planar MP, which correspond to the bending of the in-plane bonds, as well as by modes that involve the central metal. Experimentally, it has been found that an intense band at ~ 1360 cm'l, for pyrrole-substituted MP5, is sensitive to the metal oxidation state. This band, which is commonly referred to as the "oxidation-state marker, V4" corresponds to the breathing mode of C-N bonds. Shifts to higher frequency in this line result from lower it- electron density back-donation to the porphyrin eg(1t*) antibonding orbitals (e.g., FeIII -) FeIV).58 Essentially, all the porphyrin skeletal modes in the high-frequency region (1450 - 1700 cm'l) show a negative 35 linear correlation with core size for planar porphyrins.16b Roughly, a decrease in frequency of 1 cm"1 in a core size line represent an increase in core size of ~ 0.002A.56a Certain porphyrin modes in this region show sensitivity to the metal spin-state. This is especially true of the band at ~ 1580 cm'1 (v19, CaCm) which is relatively insensitive to the metal oxidation state. Other spin-state marker bands appear at ~ 1490 (V3, CaCm) and ~ 1630 cm"1 (v10, CaCm). For example, in high-spin iron(III) porphyrin complex, electrons populate the antibonding orbitals (e.g., dxz- y2) and the lengthened bonds are accompanied by the expansion of the porphyrin core, followed by the decrease in the force constants of roughly 75% of the Raman active vibrational modes.58 Doming and ruffling of the porphyrin ring also alters these frequencies, but their influence is less important than the dominant core size effect. Upon removal of an electron from the highest filled porphyrin 1t- bonding orbitals, am or am, we find that the frequencies of the stretching modes with predominantly Cbe character increase, whereas those with CaCm and CaN character decrease in the RR spectra of the cation radicals relative to the neutral metalloporphyrins, Figure 8. These structural trends seem to be essentially insensitive to 2A2u vs. 2A1u radical designation, at least for the cation radicals studied here, Figure 10.59 An Overview of this Thesis For a complete and systematic investigation of redox, vibrational, electronic and structural properties of metalloporphyrin n-cation radicals, we begin by addressing the oxidation products of cobalt porphyrins in Chapter 2. The versatility of cobalt(II) porphyrin system 36 2o. - .3033 39 - 1.0.0. 32- cane? 5.2 22- C: t C .2 0.3»? 0.6. zone . «Ow. II. 1.0.3 as... e nu!l .0». I300 lSOO I700 “00 900 RAMAN SHIFT (cm") RR spectra of (a) CuOEP; (b) CuOEP+'ClO4'. CHzClz bands are marked with an ’. CW laser power 20-35 mW. Figure 10. 37 allows us to compare the RR spectral characteristics of metal- vs. porphyrin- centered oxidized species all with the same metalloporphyrin complex. Additionally, cobalt spin-state remains constant upon variation in the ring oxidation and metal ligation states. Thus, the effects of ring-centered oxidation may be highlighted in the absence of any complication from changes in the spin-state, the contribution of which is significant in the RR spectra of metalloporphyrins in the high- frequency region (1450 - 1700 cm'l). The oxidation products of biologically-relevant iron porphyrins will be covered in Chapter 3. Characterization of oxidized iron porphyrins include a wider range of variables owing to several oxidation, ligation and spin states that are available to iron atom. As an appropriate extension of our metalloporphyrin rt-cation radical interest, we present, in Chapter 4, some preliminary results of the RR vibrational analysis of metallochlorin n-cation radicals. This study aims to develop a better understanding of the photosynthetic primary charge-separating event involving oxidized chlorophylls. Relative to the Raman spectra of metalloporphyrins, those of metallochlorins are considerably more complex owing to the symmetry reduction, to increased macrocycle conformational flexibility and to the likelihood of changes in normal mode composition of the chlorin macrocycle vibrations compared to those of the analogous porphyrin. Preliminary results and future perspectives on two topics that are still being pursued in our laboratories will be discussed in Chapters 5 and 6. Respectively, they are: 38 1. Novel B-pyrrolic substitution reactions of tetraperfluorophenyl porphyrin and 2. Metal-axial ligand vibrations in metalloporphyrins, their n-cation radicals and other related systems - a preliminary vibrational study in the low-frequency region (< 1,000 cm‘l). these topics will be presented in this chapter. 10. 11. 12. 13. 14. 15. REFERENCES Reid, L. S.; Mauk, M. R.; Mauk, A. G., J. Am. Chem. Soc..1984, 182. Suslick, K. S.; Reinert, T. J., J. Chem. Ed. 1985, §_2_, 974. (a) Dawson, J. H.; Sono, M. Chem. Rev. 1987, 1255. (b) Hewson, W. D.; Hager, L. P. In "The Porphyrins," Vol. 7, Delphin, D. Ed., Academic Press, New York, 1979, Chapter 6. Renger, G., Angew. Chem. Int. Ed. Engl. 1987, g, 643. Chang, C. K.; Fajer, J., J. Am. Chem. Soc. 1980, 1%, 848. Johnson, E. C.; Neim, T.; Dolphin, D., Can. J. Chem. 1978, 5_6, 1381. Frew, J. E.; Jones, R, Adv. Inorg. Bioinorg. Mech. 1984, 3, 175. Hanson, L. K.; Chang, C. K.; Davis, M. 8.; Fajer, J., J. Am. Chem. Soc. 1981, 1%, 663. (a) O'Malley, P. J.; Babcock, G. T., Proc. Nat. Acad. Sci. USA 1984, 81, 1098. 0)) Chang, C. K.; Kuo, M. -S., J. Am. Chem. Soc. 1979, 1Q, 3413. Fuhrhop, J. -H.; Mauzerall, D., J. Am. Chem. Soc. 1969, 9_1_, 4174. Wolberg, A.; Manassen, J., J. Am. Chem. Soc. 1970, g, 2982. Hambright, P.; Bearden, A. In "Porphyrins and Metalloporphyrins," Smith, K. M., Ed., Elsevier: New York, 1975, Chapter 12. Phillippi, M. A.; Goff, H. M., J. Am. Chem. Soc. 1982, Q4, 6026. Commoner, B.; Townsend, J.; Pake, G. Nature 1954, 174, 689. Gibson, J. F.; Ingram, D. J. E., Nature 1956, 1_78_, 871. 39 16. 17. 18. 19. 20. 21. 24. 25. 26. 28. 29. 30. 31. 32. 40 (a) Dolphin, D.; Muljiani, Z.; Rousseau, K.; Borg, D. C.; Fajer, J.; Felton, R. H., Ann. N .Y. Acad. Sci. 1973, 206, 177. (b) Spaulding, L. D.; Eller, D. G.; Bertrand, J. A.; Felton, R. H., J. Am. Chem. Soc. 1974, 9_6_, 982. Gouterman, M., J. Mol. Spectrosc. 1961, 6, 138. Edwards, W. D.; Zerner, M. Q, Can. J. Chem. 1985, 6_3_, 1763. Fajer, J.; Davis, M. S. In "The Porphyrins," Vol. 4, Dolphin, D., Ed., Academic Press: New York, 1979, Chapter 4. Fajer, J.; Borg, D. C.; Ferman, A.; Felton, R. H.; Vegh, L.; Dolphin, D., Ann. N .Y. Acad. Sci. 1973, 2_06, 349. (a) Carnieri, N .; Harriman, A., Inorg. Chim. Acta 1982, 6_2_, 103. (b) Fajer, J.; Borg, D. C.; Forman, A.; Dolphin, D.; Felton, R. H., J. Am. Chem. Soc. 1970, 9;, 3451. Gouterman, M., J. Chem. Phys. 1959, 3_0, 1139. Castro, C. E. In "The Porphyrins," Vol. 5, Dolphin, 0., Ed., Academic Press: New York, 1978, Chapter 1. Balch, A. L.; Latos-Grazynski, L.; Renner, M. W., J. Am. Chem. Soc. 1985, 1_07, 2983. Reid, L. S.; Lim, A. R; Mauk, A. G., J. Am. Chem. Soc. 1986, £3.34 8197. Fuhrhop, J. -I-I., Struc. Bonding (Berlin) 1974, E, l. (a) Brown, G. M.; Hopf, F. R.; Meyer, T. J.; Whitten, D. G., J. Am. Chem. Soc., 1975, Z, 5385. (b) Balch, A. L.; Renner, M. W., J. Am. Chem. Soc. 1986, 108, 2603. Spiro, T. G.; Stong, J. D.; Stein, P., J. Am. Chem. Soc. 1979, 101, 2648. Teraoka, J.; Kitagawa, T., J. Phys. Chem. 1980, 84, 1928. Dolphin, D.; Felton, R. H.; Borg, D. C.; Fajer, J., J. Am. Chem. Soc. 1979, 9_2, 743. Kadish, K. M.; Lin, X. Q.; Han, B. C., Inorg. Chem. 1987, 26, 4161. Salehi, A.; Oertling, W. A.; Babcock, G. T.; Chang, C. K., J. Am. Chem. Soc. 1986, 198, 5630. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 45. 46. 47. 48. 49. 50. 41 Dolphin, D.; Niem, t.; Felton, R. H.; Fujita, I., J. Am. Chem. Soc. 1975, 9_7, 5290. Mashiko, T.; Kastner, M. E.; Spartalian, K.; Scheidt, W. R.; Reed, C. A., J. Am. Chem. Soc. 1978, mg, 6354. (a) Huang, Y. -P.; Kassner, R. J., J. Am. Chem. Soc. 1979, 10_1, 5807. (b) Scheidt, W. R.; Geiger, D. K.; Haller, K. J., J. Am. Chem. Soc. 1982, _1_0_4, 495. Hill, A. O.; Skyte, P. D.; Buchler, J. W.; Leuken, H.; Tonn, M.; Gregson, A. K.; Pellizer, G., J. C. S. Chem. Comm. 1979, m. Spellane, P. J.; Gouterman, M.; Antipas, A.; Kim, S.; Liu, Y. C., Inorg. Chem. 1980, E, 386. Perrin, M. H.; Gouterman, M.; Perrin, C. L., J. Chem. Phys. 1969, 59, 4137. Godziela, G. M.; Goff, H. M., J. Am. Chem. Soc. 1986, 1% 2237. Kadish, K.; Davis, D. G.; Fuhrhop, J. H., Angew. Chem. Int. Ed. Engl. 1972, fl, 1014. Ichimori, K.; Ohya-Nishiguchi, H.; Hirota, N .; Yamamoto, K., Bull. Chem. Soc. JPn. 1985, 58, 623. Konishi, S.; Hoshino, M.; Imamura, M., J. Am. Chem. Soc. 1982, 104, 2057. Hambright, W. P.; Thorpe, A. N .; Alexander, C. C., J. Inorg. Nucl. Chem. 1968, Q, 3139. Evans, D. F., J. Chem. Soc. 1959, 2003. Erler, B. S.; Scholz, W. F.; Lee, Y. J.; Scheidt, W. R.; Reed, C. A., J. Am. Chem. Soc. 1987, 102, 2644. Scholz, W. F.; Reed, C. A.; Lee, Y. J.; Scheidt, W. R.; Lang, G., J. Am. Chem. Soc. 1982, _1_04, 6791. Kadish, K. M., Prog. Inorg. Chem. 1986, 3_4, 435. McConnell, H. M., Chestnut, D. B., J. Chem. Phys. 1958, _2__8_, 107. Phillippi, M. A.; Goff, H. M., J. Am. Chem. Soc. 1982, _1_0;4, 6026. Goff, H. M.; Phillippi, M. A., I- Am. Chem. Soc. 1983, 10_5, 7567. 51. 52. 53. 54. 55. 56. 57. 58. 59. 42 Morishima, I.; Takamuki, Y.; Shiro, Y., J. Am. Chem. Soc. 1984, 1%, 7666. Shimomura, E. T.; Phillippi, M. A.; Goff, H. M., J. Am. Chem. Soc. 1981, fl, 6778. Spreer, L. O.; Maliyackel, A. C.; Holbrook, S.; Otvos, J. W.; Calvin, M., J. Am. Chem. Soc. 1986, 1%, 1949. Hinman, A. 5.; Pavelich, B. J.; McGarty, K., Can. J. Chem., submitted. Itoh, K.; Nakahasi, K.; Toeda, H., J. Phys. Chem. 1988, 22, 1464. (a) Parthasarathi, N .; Hansen, C.; Yamaguchi, s.; Spiro, T. g., J. Am. Chem. Soc. 1987, 192, 3865. (b) Kitagawa, T.; Ozaki, Y., Struct. Bonding (Berlin) 1987, 6_4, 71. Andersson, L. A.; Loehr, T. M.; Chang, C. K.; Mauk, A. G., J. Am. Chem. Soc. 1985, 192, 182. Spiro, T. G., Israel J. Chem. 1981, 2_1, 81. Oertling, W. A.; Salehi, A.; Chung, Y. C.; Leroi, G. E; Chang, C. K.; Babcock, G. T., J. Phys. Chem. 1987, fl, 5887. CHAPTER II PREPARATION METHODS AND SPECTROSCOPIC PROPERTIES OF 0x11312151) COBALT PORPHYRINS In contrast to iron and copper, which dominate the scene of transition metal biochemistry and are components of a variety of metalloproteins, cobalt occupies a relatively modest position in biology. Although Co(II) porphyrins do not occur naturally, the related vitamin B12 corrin system catalyzes molecular rearrangements via the Co(II) oxidation state in at least some of its enzymatic reactions.1 Most of the literature on cobalt(II) in biochemistry concerns its effects in various metal-activated enzyme systems.2 Many of the typical Mg2+, Mn2+, Fe2+ and Zn2+ activated metalloenzymes can work with Co2+ at a reasonable rate. A general observation seems to be that cobalt generates catalytic activity, often approaching, and sometimes exceeding that of the native enzyme.3 Thus, owing to its spectroscopic and magnetic properties, substitution of Co2+ for the native metal is considered to be a useful probe of the metalloenzyme active site. The utility of the Coz+ ion is also well-documented in hemoproteins and their synthetic analogs in which cobalt substitution has provided additional information for a better under- standing of spectroscopic, structural and reactivity properties of these metalloproteins. For example, Co(II)-reconstituted hemoglobins and myo- globin have been extensively investigated in an attempt to obtain further 43 44 information on heme-heme interations.4 Hoffman and coworkers have shown that cobalt hemoglobin displays a reversible and cooperative uptake of oxygen that is qualitatively similar to that of hemoglobin.5 Cobalt porphyrins have also been applied in disproportionation reactions of hydrogen peroxideé, which in vivo is prompted by catalase, as "shift reagents" to determine the stereochemistry of molecules with which they associate7a, and as the catalyst for various redox reactions of the axial ligands.7b The study of the oxidation products of cobalt porphyrins is also of great interest. The versatility of the cobalt porphyrin system allows us to compare metal vs. porphyrin centered oxidation products for the one-electron case, and 2A2u vs. 2A1u radicals for the two-electron case, all with the same metalloporphyrin species. The latter species are of principal interest owing to their relevance to the Compound I type enzyme intermediates.8 Addi- tionally, cobalt exhibits only one spin-state (S = 1/ 2) and this can avoid complications from changes in spin-state upon oxidation, a phenomenon observed in iron porphyrins.9 It is the aim of this chapter to present first the preparation of oxidized cobalt porphyrins, followed by their spectroscopic characterization by means of UV-vis, ESR and magnetic susceptibility, IR and particularly RR technique to highlight the vibrational properties of porphyrin n-cation radicals. As noted previously, the pyrrole-substituted porphyrins; i.e., OEP, will be utilized in this study. 1. Preparative Methods. OEP was synthesized according to the published method.10 The meso-tetradeuteration of OEP was carried out in D2504- D20 as described by Bonnett and coworkers.11 CH2C12, freshly distilled from CaHz, was used as a solvent for all spectroscopic studies unless otherwise specified. 45 CoHOEP 1: To a refluxing chloroform solution of HZOEP was added a three-fold molar excess of cobaltous chloride-sodium acetate (1:1) dissolved in minimum amount of methanol. The progress of the reaction was followed spectrophotometrically until quantitative metallation was achieved. The crude reaction mixture was first evaporated to dryness. The residue was redissolved in minimum amount of dichloromethane and was washed successively with dilute acid (once) and with water (several times), to remove excess inorganic salts, and dried over anhydrous sodium sulfate. The dichloromethane solution was evaporated to dryness and the resulting fluffy CoOEP(1) solids were collected. Alternatively, the Co(II) derivative may readily be obtained from the reaction of free base porphyrin with cobaltous acetate (10X molar excess) in refluxing acetic acid12 with a similar workup procedure as described above. COHIOEPX‘ 2: Method 1. For x- = Br,‘ CoHOEP (30 mg) was suspended in methanol (30 ml) containing ~ 3 ml of 48% hydrobromic acid.13 When the suspension was stirred at room temperature for several hours, the solution gradually changed to a deep red color, until the red crystals were precipitated. The resulting crystalline precipitates where collected, washed subsequently with water, methanol, petroleum ether and dried at room temperature. Yield: > 90%. Method 2. CoHOEP (50 mg) was dissolved in 10 ml of chloro- form, and to this solution was added 10 ml of methanol and ~ 0.2 ml of 48% hydrobromic acid.14 The mixture was stirred at room temperature for 2 h. and then an additional 10 ml of chloroform was added. The reaction mixture was washed with water, dried over anhydrous sodium sulfate and evaporated to dryness. Yield: > 90%. The iodo- and chloro- derivatives may also be prepared as above by using an appropriate concentration of 57% hydriodic and 37% hydrochloric acid, respectively. CoHOEP+~X' 3: This species was obtained by stirring a dry dichloromethane solution of 1 with a three-fold molar excess of solid anhydrous silver perchlorate at room temperature for about 1 h.15 The bright red solution of the neutral CoHOEP turns brownish-red upon completion. The solution is then filtered and the product can be isolated by precipitation with hexane. The cobaltous porphyrin cation radical is stable only in the presence weakly coordinating ligands such as ClO4", BF4‘ and PF6" but not Br' where the formation of bromo- cobaltic derivative is favored. CoIHOEP‘l'QX' 4: When X' = Br', a dichloromethane solution of 1 may be oxidized by means of molecular bromine. The first step, which requires ~ 0.5 moles of bromine, brings about the oxidation of 1 to 2. Further oxidation of this trivalent complex, using an additional ~ 0.5 moles of bromine gives the green cationic radical species 4. When X- = ClO4", this dibromide species dissolved in dichloromethane is treated with silver perchlorate16a or, alternatively, a direct two- electron oxidation of CoHOEP may be achieved in dichloromethane by using a large excess of ferric perchlorate salt.16b Com(ROH)ZOEPX’ 5: When R = CH3, this complex was obtained by oxidation of 1 with HX(X‘ = ClO4',Br‘) in dichloromethane- methanol 47 (3:1). The diaquo adduct may be prepared in the same manner. Weakly coordinated methanol and water molecules are readily placed with stronger ligands such as pyridine.17 4 may also be prepared indirectly by titrative addition of methanol or water to a dichloromethane solution of CoIIOEP+°X', ComOEPX‘, and CoHIOEP+-2X' in the order of increasing the required volume of ROH for complete conversion.18 Instrumentation: UV-vis spectra were recorded by using Shimadzu UV-160 and Cary 219 spectrophotometers. The ESR spectra were measured on a Varian E-4 spectrometer. The g-value calibration and spin quantitation were obtained by using Fremy's Saltlsarb Magnetic susceptibility was obtained on an SHE SQUID susceptometer at a field of SkG. Infrared spectra of Co OEP compounds were examined as KBr disks by using a Perkin-Elmer 599-IR spectrophotometer. Raman Spectra were measured with a Spex 1877 Triplemate and OMA H electronics (Oertling et al., J. Phys. Chem. 1987, 9_1_, 5887-5898). Laser emission at 363.8 nm was provided by a Coherent Innova 90-5. Argon Ion Laser. 2. Electronic Absorption Spectra: As described by Johnson”, the various cobalt porphyrin complexes can be differentiated by their visible spectra and changes can be followed conveniently by using a spectrophotometer. The general pattern of a normal pyrrole-substituted metallopor- phyrin spectra is well known.20 A dichloromethane solution of 1 exhibits such a pattern with two moderately intense bands, the so-called oz and B bands (8 ~ 104 M'lCm'l) at 551 and 517 nm, respectively, and the strong Soret band at 391 nm (8 ~ 105 M‘1Cm'1). Of the two visible bands, the a band has the higher intensity, Figure 1. . Formation of the Co(III) complex is accompanied by the red-shift of both Soret and visible peaks with respect to those in Co(II), a phenomenon pointed out by Corwin and coworkers. They suggested that this is a general trend for square planar porphyrins bonded to ligands in the octaehedral positions. The electronic spectrum is always "normal", but the ratios of the visible on and [3 bands, as well as the exact position of the Soret band are solvent-dependent to an exceptional degree, Figure 1. A more recent study by Wang and Hoffman21 support the earlier results but also demonstrates that the perturbations by the ligand not only shift the overall spectrum but also systematically change both the frequency difference between the Soret and the visible bands and the relative intensity of tea/EB. Thus, the position of the a band shifts less rapidly than that of the Soret band and the individual axial ligands affect the spectra in the following order of increasing red shift and decreasing Ea/EB ratios: "e""(phantom ligand) < H20 ._._m2m:.ZH 23243. bid, i300 I500 I700 RAMAN SHIFT (cm") H00 900 Figure 8. RR spectra of cobaltous octaethylporphyrin in CHzClz. 62 vibrations. The weak mode at 1476 cm'1 may correspond to V28 (CaCm), also reported for NiPP (PP = protoporphyrin-IX) at 1482 cm-1.32 The intense polarized oxidation-state marker band (V4) appears at 1379 cm'1 in Figure 8a and is predominantly of CaN stretching character. The depolarization ratio, the absence of a deuterium shift, Figure 8b, and the clear spectral analogy to the other well-characterized metalloporphyrins“: 32 forms the basis of this assignment. The effects of metal-centered, one-electron oxidation accompanied by axial metal ligation are demonstrated by comparing the spectra of CoHOEP and CoIH(MeOH)ZOEPClO4' depicted in Figure 9a,b. The oxidation-state marker, V4, increases from 1379 to 1383 cm'l, reflecting depopulation of the porphyrin it" orbitals caused by metal oxidation.33 There is little systematic change in frequency of modes above 1450 cm'l, indicating that the core size of the porphyrin ring does not change significantly upon metal oxidation and addition of axial methanol ligands-32:34 Figure 9c, (1, shows the spectra of the cobaltous and cobaltic OEP+° complexes. Neglecting differences in relative intensity produced by the differences in Soret absorption, the RR spectra of 3 and 4 are essentially identical and (above 1300 cm’l) radically different from those of the neutral ring compounds in Figure 9a, b. We assign the V3, V11, V2 and V10 frequencies of CoIIOEP+'ClO4', for example, at 1505, 1604, 1620, and 1642 cm‘l, respectively, the assignments of which are based upon similarity in depolarization ratio measurements and isotope data with the authentic 7t- cation radical ComOEP+-2CIO4', Figures 9c, d and 10. 63 'i379 )‘ex - 363.8 nrn oi Co" OEP "' i590 — i375 " 00" - 05.2 -t303 bl Co'iMeOHi 206900: (0 Q - 2'05 - - 0 - I643 i000 RAMAN INTENSITY c) Co. 0E P" Clo; I041 - 03“ "' I505 - .54; i6” dl Co" OEP” 2cm; A g A- A. L Asoo I100 I500 1500 A I7oo RAMAN SHIFT Icm") Figure 9. RR spectra excited at 363.8 nm (- 35 mW) of C005? and its oxidation products. (a) CoOEP; (b) ComMeOI-D20EPC104‘; (c) ConOEP+ClO4’; (d) Comosr+-2c104-; (e) Com(MeOI-I)20F.P8r'; (0 CoanEPBr‘; (g) ComOEP+°28r‘. Solvent, dry CH2C12 except (b) and (e) which contain ~ 5% methanol. Solvent bands are marked with an ’ 64 o ‘2 I X... - 3535 nm e) Co"(MeOHizOEP Br' N O ‘3 t _. i378 fl Co'OEP Br' RAMAN INTENSITY g .780 900 000 1300 A 1500 RAMAN SHIFT (cm") Figure 10. Electronic absorption spectra of oxidation products of CoOEP. (a) Com(MeOH)20EPBr'(-); (b) ComOEPBr' (—): (c) ComOEP+-25r' (---). the small feature at 401 nm is due to 1% H40EP2+23r‘ contamination as discussed in the text. Solvent, dry CHzClz, except for (a) which contains ~5% methanol (MeOH). 65 The R spectra excited at 363.8 nm of the bromide adducts of oxidized CoOEP are shown in Figure 11. The R spectrum of CoHI(MeOH)zOEPBr', Figure 11a, is similar to that of CoOEP and Com(MeOH)zOEPClO4', Figure 9a, b. The R spectrum of COmOEPBr' is shown in Figure 11b. While V4 (1378 cm‘l) is slightly lower than in the other Co(III) compounds, the core-sensitive bands above 1450 cm‘l, particularly V10 (1657 cm‘l), have increased in frequency, supporting a slight core contraction in this species.32r34 The spectrum of the it-cationic radical species, ComOEP+°2Br'(Figure 11c) is dominated by v2 (1611 cm-1) and v11 (1600 cm-1) modes, similar to the other n-cations discussed earlier. There is, however, no band easily assignable to V4. The features at 1460 and 1648 cm"1 are assigned to V28 and V10, respectively. The V3 vibration, which is enhanced in Soret RR spectra of neutral metalloporphyrins, does not appear in these spectra, but a polarized feature at 1497 cm'1 present in the spectra obtained in resonance with the 670- nm transition of the dibromide cation radical is assigned to V3. (Spectrum not shown).18 Table 1 summarizes the vibrational frequencies of CoOEP derivatives. Thus, in the high-frequency region, the modes involving primarily CaCm stretching character (V3 and V10) decrease in frequency upon oxidation of the porphyrin ring, while modes involving primarily Cbe stretching character (V11 and V2) increase in frequency. The frequency of V4, primarily a CaN stretch, decreases upon formation of porphyrin rt-cation radical. These trends are common to other metalloporphyrin system (i.e., CuOEP, ZnOEP) not presented here.18 66 4 m C 2 0 P E 0 l O C )‘u - 363 8 nm e h g i a) "4 D) de >P_m2u._.ZH Z<2 :mZWFZH 242$”. '300 1500 woo woo RAMAN SHIFT tern") RR spectra of ferric chloride porphyrins in CHzClz. Figure 3. 82 >2m2uhZH Zdidm ITOO RAMAN SHIFT km") RR spectra of ferric chloride porphyrin 1c cation radicals in CHzClz Figure 4. 83 may be assigned by considering their frequency shifts relative to the neutral five- and six-coordinate ferric OEP species. Inspection of Table I shows that the pattern of frequency shifts observed for the valence +2 metalloporphyrin derivatives is reproduced when the ferric n-cation radicals are considered to be five—coordinate in solution (column 4); assuming a six-coordinate solution state for the radicals (column 5) produces poor agreement with our earlier data. Two conclusions follow from these data. First, the porphyrin core geometries of both 1 and 2 in solution are characteristic of a five-coordinate state, consistent with the suggestion that axial ligand dissociation occurs when 2 goes into solution.17 Second, the pattern of vibrational frequency changes that occurs upon ring oxidation of +2 metalloporphyrin derivatives3C appears to hold as well for the more complex, but biologically more relevant, iron porphyrins. High frequency RR scattering from Iron protoporphyrin-IX lt-cation radical: Figure 5 depicts the RR spectra of OEPFeCl and its one-electron oxidation product, OEP+°Fem(Cl')(SbCl6’). Also shown are the spectra of the analogous protoheme species (Figure Sc and d). As demonstrated earlier, the OEP compounds exhibit a five-coordinate, high-spin (S = 5/ 2) configuration for the heme iron in solution; the protoheme spectra are consistent with the same spin and coordination state assignment for these species.20 Table 2 compares the vibrational assignments of iron PP-IX with those of OEP complexes. The decrease in frequency for modes with significant Cbe character for the protoheme species relative to OEP species is expected on the basis of different pattern of peripheral substitution, also demonstrated in the spectra of the EPI complexes. The protoheme RR spectra are more complex owing to the Eu modes, V33 and V37, which occur at 1530 and ~ 1553 cm'l, respectively, in the neutral form, Figure 5021 The position of V38 in the cation radical is difficult to determine from our RR spectra excited at 351.1 nm, Figure 5d, but presumably can be measured by IR spectroscopy.3C We interpret the broad feature extending from 1570 - 1600 cm'1 in the FeHIPP-IX+° spectrum to be composed of three overlapping bands: V11 (Cbe) occurring at ~ 1570 elm-1, v2 (Cbe) at 1593 cm-1 and V37 (CaCm) at ~ 1585 cm-1. The ~ 20 cm'1 increase in the modes (V11, V2) that are predominantly composed of Cbe stretches is expected upon oxidation of ferric porphyrin lt-cation radical, Table 1. These frequency increases are partially obscured in the spectrum in Figure 5d by the frequency decreases of V37 (CaCm) in the cation (~ 1585 crn‘l) relative to the neutral (1590 cm‘l). These spectra demonstrate that the changes in the vibrational frequencies that accompany ring oxidation of these ferric compounds are consistent with those observed for other cation radical species. Thus, in general, we find that stretching modes with predominantly CaN character (V4) or CaCm character (V3 and V10) decrease in frequency, while those with (3be character (V11 and V2) increase in frequency when the porphyrin ring is oxidized, Tables 1 and 2. We see these frequency shifts regardless of whether the electron is extracted from the a1u(7c) or a2u(1t) molecular orbital. 8S >._._mZm:.ZH Z<§ 1500 cm‘l), where macrocycle stretching modes with CaCm and Cbe character occur, the MC spectra are more complex. For example, Boldt g1a_l. have identified ten vibrations in the 1450- 1700 cm'1 region of the RR spectra of t-NiOEC (solid state) by using excitation at various wavelengths.22 In the present study, we focus on two 95 c) CuMeOEC >._._m2wh2H Z<§Fttby.“ bbbbbbbb a bbbbbbb ‘ eh VG - up a“ a“. L. 2. 8M mm . . a 3.8 3 % :.3~ut .a .3 3.8. 4. a ........ a: Llatlamt ...... 4. ..... Lt,» bhttPFDFIDD ; : ...... 109 .3 E 3 8333.8“. 9.23% a: 6:0 532035020938. Co 32-—m. 2.3% «was. 82?: .8300.» 2:. N 9.sz a“ 'F" DDDDDDDDD ‘H O DDDDDDDDD q. bbbbbbbbb d DDIbDDDDD Flip} I‘D’FNDhPDLFDD bbbbbbb 3- .hbbbbhp)>.l-.-1DI.or t d.‘ 2:4... .. 5:. . . .3 ~ 3 ... nz .. . .6 a a. v..x v 31 .0 .2 g 9* +0.8. LE’fl ””””” d > .PIFL’bJ.’t’,”rb*bi‘bit”lw ........ a ....... J k v: . u. .2. a. +4.. .1... 9 1 a .98 3 M an...» I .Q g ca .98. '..”> ’0”””’p” 3. Duct. .6 r .93. In 110 trans-octaethylchlorin-d4 The tetradeuterated species was prepared by the same method as described above; however, the sample was permitted to sit at room temperature for ~ 72 hours. One hundred percent deuteration was established by the disappearance of proton NMR signals at 8.9 and 9.7 ppm due to 7,8- and a,B- protons, respectively, Figure 6d. M5: m/ e 540(M+), 511 (M‘H'), Figure 7c. trans-octaethylchlorin-dz(a,B) t-OEC-d4 (10 mg) was subject to "back-exchange" in 2.1 ml of H2504(96.4 percent):H20(6:l, v/v) followed by the same work-up as mentioned above. This method gave ~ 90 percent deuteration at 01,8- position and complete proton recovery at the 7,5- sites, Figure 6b. octaethylporphyrin-d2 To a refluxing dichloromethane solution of dideuterochlorin was added a solution of DDQ (excess) in benzene according to the published methods (20,21). The green color of the chlorin solution turned red almost instantaneously. After refluxing for an additional 0.5 h, the solution was evaporated to dryness and chromatographed on alumina with chloroform. The visible absorption spectrum and the proton N MR signal at 8 10.2 showed the characteristic OEP spectrum. M5: m/e 536(M+), 507(M+-29), and 268(M++); mp. 322-324°C. 111 CONCLUSION The original procedure reported for the tetradeuteration of OEP (80 mg in D2504zDzO, 2 ml; 9:1, w/v) has been modified to prepare three forms of selectively deuterated trans-octaethylchlorin; namely, -d2(a,B),-d2(y,5), and - d4 derivatives. This method offers an alternative approach to that of acetic acid-d1 for H/ D exchange at ambient temperature in unsymmetrically substituted tetrapyrrolic macrocycles. It is also particularly attractive for deuterium labeling in systems where prolonged refluxing conditions are not desirable. The partially-deuterated porphyrin and chlorin complexes are currently being utilized to verify the proposed "quadrant-localized" vibrational modes in the chlorin macrocycleszz, by means of IR and resonance Raman spectroscopy. 10. 11. 12. 13. REFERENCES Fajer, I., Borg, D.C., Forman, A., Dolphin, D., I. Am. Chem. Soc.,1970, 42: 3451. ., Mengersen, C., Subramanian, I., Fuhrhop, I.W., Smith, K. M., Z. N aturforsch, 1974, gag 1827. Norris, I. R., Scheer, H., Druyan, M.E., Katz, I.I., Natl. Acad. Sci. USA, 1974, _7_1: 4897. Feher, G., Hoff, A.I., Isaacson, RA., Ackerson, L.C., Ann. N. Y. Acad. Sci., 1975,1416: 239. Morishima, I., Takamuki, Y., Shiro, Y., I. Am. Chem. Soc., 1984, 126: 7666. Godziela, G.M., Goff, H.M., I. Am. Chem. Soc., 1986, 1%: 2237. Urban, M. W., N akamoto, K., Kincaid, J., Inorg. Chim. Acta, 1982, 61: 77. Kincaid, I. R., Urban, M. W., Watanabe, T., Nakamoto, K., I. Phys. Chem., 1983, 82: 3096. Ozaki, Y., Kitagawa, T., Ogoshi, H., Inorg. Chem., 1979, 1_8_: 1772. Andersson, L. A., Sotiriou, C., Chang, C. K, Loehr, T. M., I. Am. Chem. Soc., 1987,1025 258. Bonnett, R., Stephenson, G.F., Proc. Chem. Soc., 1964, 291. Bonnett, R; Gale, I. A. D., Stephenson, G. F., I. Chem. Soc. (c), 1967, 1168. Smith, K. M., Langry, K. C., deRopp, I. 5., I. Chem. Soc. Chem. Comm., 1979, 1001. 112 14. 15. 16. 17. 18. 19. 20. 21. 113 Cavaleiro, I. A. 5., Smith, K. M., I. Chem. Soc., Perkin Trans. 1973, 1, 2149. Hickman, D.L., Goff, H.M., I. Am. Chem. Soc., 1984, Q6: 5013. Woodward, R. B., Skaric, V., I. Am. Chem. Soc., 1961, 88: 4676. Stolzenberg, A. M., Laliberte, M. A., I. Org. Chem., 1987, 5_2: 1022. Wang, C. B., Chang, C. K., Synthesis, 1979, 548. Whitlock, Ir., H. W., Hanauer, R., Oester, M. Y., Bower, B. K., I. Am. Chem. Soc., 1969, 21: 7485. Eisner, V., Linstead, R P., I., 1955, 3749. Barnett, G. H., Hudson, M. F., Smith, K. M., Tet. Lett., 1973, 2887. Boldt, N. I., Donohoe, R. I., Birge, R. R., Bocian, D., I. Am. Chem. Soc., 1987,fl2: 2284. CHAPTER V NOVEL B—PYRROLIC SUBSTITUTION REACTIONS OF TETRAPERFLUOROPHENYLPORPHYRIN Tetraperfluorophenylporphyrin (TFPP) is an electron-deficient porphyrin having highly electron-withdrawing yet chemically inert meso- perfluorophenyl substituents. This pattern of peripheral substitution is expected to raise the redox potentials of the porphyrin macrocycle dramatically. For example, half-wave potentials of TFPPFeCl are positively shifted by 0.4 and 0.32 V with respect to those of TPPFeCl for the first and second oxidation waves, respectively.1 On the other hand, flourine substitution, particularly on phenyl C-2 and C-6 positions, provide the porphyrin ring with added steric protection against oxidative attack which is commonly directed at the meso carbons. This reaction often results in the cleavage of the ring and formation of an open-chain tetrapyrrolic structure.2 This type of electrophilic substitution is operative in biological heme catabolism and it has also been demonstrated with the OEP and TPP model compounds.3 Thus, increased stability against self-destruction as well as higher oxidation potentials make the TFPP system a powerful oxidation catalyst for the study of cytochrome P-450 mediated reactions.4 For example, Chang and Ebina have shown that the TFPPFeCl-catalyzed hydroxylation of cyclohexane to cyclohexanol proceeds in 71% yield compared with the 5% 114 115 yield obtained from TPPFeCl.5 Apparently, the electron-deficient TFPP com- plex forms a' more reactive and electrophilic oxene complex (TFPP+°FeIV=0). During a chemical and spectroscopic investigation of the metallo TFPP complexes, Gouterman and coworkers observed a new series of TFPP adducts. These adducts were formed when the free base was treated with AgNO3 in acetic acid or, alternatively, the corresponding metal derivatives (e.g., Ag2+, N221; Zn2+ and Cu2+) were dissolved in acetic acid in the presence of catalytic amount of concentrated nitric acid, under refluxing condition. A This work aims to elucidate the structures of these TFPP, derivatives by Uv-vis, FAB-M5 and lH-NMR studies. General Characteristics of TFPP a. Reduced Affinity to Metal Ions: The insertion of Coz+ ion, for example, into HzTPP may be carried out by refluxing a chloroform-methanol solution of the free base containing cobaltous chloride (or acetate), while under identical condition, most of H2TFPP remains as a free base even after 24 hours. Thus, CoTFPP may be prepared by refluxing an acetic acid solution of H2TFPP which contains a large molar excess (10x) of cobaltous acetate. b. Resistance Toward Acid Demetallation: N iTFPP, for example, resists demetallation considerably. N iTPP can be completely demetallated with H2504 at RT, while a super acid (e.g., FSO3H) is required for the TFPP derivative. Gouterman e_t 81. reported an alternative method by refluxing a trifluoro acetic acid solution of N iTFPP for ~ 72 hours.14 c. Reduced Proton Affinity: H2TFPP exists as free base in acetic acid. In a marked contrast, HZOEP exists as a monocation in acetic acid.6 The great dicationic salt of H2TFPP may readily be formed by the addition of TFA, 116 HN O3 and HClO4 to a dichloromethane solution of the free base. Table I summarizes the optical changes associated with the formation of H4TFPP2+2X'. Table I. Absorption Spectra (nm) of Dicationic Salt of H2TFPP in CHzClz. Compound Soret IV III II I H2TFPP 412 506 - 583 655 H4TFPP2+°2X' X' = TFA‘ 430 538 574 625 N 03' 434 543 579 629 ClO4' 433 ? 580 631 The preceding observations as well as the increase in oxidation potentials of TFPPFeCl by ~ 0.4V with respect to those of TPPFeCl demon- strate that introduction of the fluoro groups on the phenyl rings results in a significant reduction in the electron density on pyrrolic nitrogens. d. Optical Absorption Properties: TFPP complexes feature unusual electronic spectra which separate them from other tetraphenyl porphyrins.7 The energies of absorption maxima in the free base and the metal derivatives fall between those of OEP at higher energy and TPP at lower energy, Table 2. The most remarkable difference is observed in the spectrum of H2TFPP in which an absorption maximum near 550 nm (band 111) is absent, in contrast with that of HZTPP.7 117 Table 2. Comparison of H2TFPP Absorption Maxima (nm) in CH2C12 with Those of HZTPP and HZOEP. Compound Soret IV III II I H2TFPP 412 506 - 583 655 HZOEP 397 496 531 565 615 Kim and coworkers later confirmed the phenomenological cause of this anomalous behavior.8 They found that electron-withdrawing substituents in the phenyl ortho position cause the diminishing of the extinction coefficients of bands 1, HI, and Soret; whereas, bands II and IV are not affected. Moreover, phenyl substitution in other positions have minimal effect. The visible absorption spectra of TFPP complexes exhibit strong solvent-dependent maxima shifts, Table 3. Thus, if optical changes, which are accompanied by TFPP substitution reactions, are to be used to partially characterize the newly-formed adducts, the solvent-induced shifts must also be taken into consideration. 118 Table 3. Solvent-induced Absorption Maxima Shifts (nm) of TFPP Adducts. Compound Solvent Soret IV 111 II I H2TFPP DMF 410 504 — 579 654 CH2C12 412 506 - 583 655 0H 417 508 - 586 659 ¢CH3 417 509 - 586 658 N iTFPP CH2C12 404 524, 558 620 ¢CH3 407 526, 559 621 PY 429 553 621 CuTFPP CHzClz 408 534, 570 ¢CH3 416 539, 573 ZnTFPP CHzClz 413 543, 577 606 ¢CH3 421 547, 580 607 e. 1H-NMR: H2TFPP and its N12+ and th+ complexes feature a single peak at 8.50, 8.77 and 8.99 ppm, respectively, in CDC13 characteristic of B-pyrrole protons. The upfield shifts of the B-hydrogen peaks in the TFPP adducts as compared with those observed in the TPP system (e.g., HZTPP, H3 = 8.75 ppm) has been explained in terms of the degree of buckling of the porphyrin nucleus due to meso-fluorophenyl substituents.9 Apparently, the steric interaction between the fluoro groups and the B-hydrogens of the pyrrole rings results in increased c0planarity of the phenyl groups with the porphyrin ring in the TFPP system. This, in turn, reduces buckling of the macrocycle and increases the ring current effect. 119 f. Reversible Oxidation Reactions of MTFPP Derivatives - Electronic Absorption Characteristics: Ferric perchlorate salt, Fe(ClO4)3°XHzO, serves as a chemical oxidant of general utility and provides adequate redox potential for at least single-electron oxidation of all common porphyrins (81 /2 = 1.2V vs. SCE). The utility of this oxidant has been demonstrated in several instances in the previous chapters. However, it falls short of oxidizing Cu2+ and N 12+ complexes of TFPP. The Zn2+ adduct demetallates immediately upon oxidation. This may result from 1) large ionic radius of Zn2+ ion, 2) electron- deficient porphyrine ring and 3) further weakening of the metal-porphyrin bonds upon removal of an additional electron from the porphyrin )t-system to form a cation radical. However, the Ag2+ and Co2+ complexes of this porphyrin are readily oxidized at the metal to form the corresponding valance +3 derivatives. The absorption-shift pattern observed for both systems upon oxidation agrees well with the Ag(III) and Co(III) complexes of other porphyrins, Table 4.10: 11 Table 4. Changes in the Optical Absorption (nm) in CHzClz of the AgII and CoII Complexes of TFPP Upon One-Electron Metal-Centered Oxidation. Compound Soret Visible Bands AgHTFPP 419 537, 570 AngFPPClO4' 418 527, 559 CoHTFPP 404 527 ComTFPP(HzO)2ClO4‘ 422 536 w m 120 Our attempt to chemically oxidize CoIITFPP by more than one electron resulted in the partial formation of CoIIITFPP+°2ClO4' (e.g., 666 nm) and that the CoIII species (e.g., 422 and 536 nm) persisted for the most part. The presence of the radical was conformed by observing a broad, residual ESR signal at room temperature. Qualitatively, it is expected that the TFPP radicals are more prone than their TPP counterparts to adopt 2Aw electronic ground state or that the fluorophenyl groups may completely stabilize the a2U orbital relative to the an} orbital by withdrawing electron density from the meso carbons.12 The ZAIU vs. 2A2U state of the MTFPP radicals may be crucial for selective peripheral functionalization of the porphyrin ring either at the B-pyrrolic or on the bridging meso-carbon positions, as was recently reported by Catalano and coworkers for the analogous TPP complexes.13 EXPERIMENTAL The reaction conditions applied in this study may be divided into three categories as follows: 1. The free base H2TFPP was treated in glacial acetic acid with a large excess (e.g., 10x) of metal nitrate salt (e.g., Ag2+, Cu2+) under refluxing condition for ~ 6 hours. 2. The desired MTFPP (e.g., Ag2+, N121“) was initially prepared from refluxing an acetic acid solution of the free base with a large molar excess (e.g., 10x) of metal acetate. Subsequently, MTFPP was treated with acetic acid in the presence of catalytic amount of concentrated nitric acid under refluxing condition for ~ 6 hours. 3. The free base was refluxed in acetic acid in the presence of a large molar excess of metal acetate (e.g., Zn2+) and catalytic amount of concentrated nitric acid for ~ 6 hours. 121 In all above cases, removal of the solvent gave a green solid which was redissolved in dichloromethane and chromatographed on silica gel TLC plate (dichloromethane/hexane 9:1). The individual fractions in which their optical spectra contained two bands in the 500-600 nm region and displayed no fluorescence under the UV light were subsequently demetallated in TFA under refluxing condition for 48 - 72 hours. The TFA solution was evaporated to dryness and the solid was taken up in dichloromethane and rechromatographed on silica gel (dichloromethane/ hexane 9:1) to afford the free base adducts. This step, in some instances, resulted in the separation of two or more fractions, which are not separable when they are metallated. Moreover, demetallation would facilitate the 1H-N MR analysis of paramagnetic species (e. g., Cu2+, Ag2+). Results Generally, three types of products were obtained based on their optical absorption characteristics: a. Type 1 features two bands of comparable intensity in the 500 - 600 nm region, characteristic of metalloporphyrins and designated as M(2-B). Although Gouterman _t _1.7 have demonstrated the solvent-dependent intensity variation of MTFPP visible bands, Figures 1 and 2, the so-called a band intensifies drastically in M(2-B) species, Figure 3. This figure also demonstrates that Cu(2—B) spectrum is red-shifted relative to CuTFPP. Type 1 species may be prepared by refluxing an acetic acid-nitric acid (catalytic, ~ 3 drops) solution of MTFPP for ~ 6 hours. b. Type 2 species also displays a two-banded visible region; however, in the absence of the central metal. This spectrum may arise from the mono- 122 Macaw-u Wavelength (nm) Figure 1. Absorption spectra of tetrakis(perfluorophenyl)porphyrin complexes taken in CHzClz at room temperature: (A) free base; (B) Zn complex, (C) Cu complex; (D) Pd complex. (From Ref. 7.) 123 Figure 2. Solvent-dependent absorption spectra of zinc tetrakis(perfluoro- phenyl)porphyrin in visible region: (A) dry toluene: (B) toluene shaken with HzO;(C) 75% toluene and 7.5% CHgCN; (D) 75% toluene and 25% ethanol; (E) 75% toluene and 25% pyridine: (F) 75% toluene and 25% triethylamine. (From Ref. 7.) 124 4|6 4O ABSORPTION \.___ -m 600 m“ Figure 3. Electronic absorption spectra of CuTFPP(-) and Cu(2-B) in CHzClz. 125 and diprotonated free base of the functionalized TFPP adduct, designated as HM(2-B). In the absence of the exact identity of peripheral groups, type 2 may be distinguished from type 1 by their active emission properties.” Type 2 adducts are formed as the by-products of M(NO3)2/HOAC or HOAC/HNO3 reaction conditions. c. Type 3 adduct, designated as HM (4-B), shows a free base-like spectrum in which band 111 is present, in contrast with that of H2TFPP. The spectrum is also red-shifted relative to H2TFPP. The species derived from AgTFPP features a band I which is more intense than bands 11 and III, Figure 4. Type 3 species may be prepared when M(2—B) is demetallated in TFA or, an acetic acid-nitric acid (catalytic, 5 drops) solution of MT FPP is refluxed for ~6 hours. The splitting of the exo B-pyrrolic protons in the 1H-NMR spectra of the newly-formed TFPP adducts provides evidence for the functionalization of B-pyrrolic double bonds. The extent of substitution is not clear in all cases; however, resolution of nine or ten bonds in a few instances reveal that up to two or three B-pyrrolic substituents may be present, Figure 5a and b. In the FAB mass spectra of these adducts, observation of peaks at m/z 975 (M+, H2TFPP), 993 (M+, H2TFPP + H20), 1006 (M+, H2TFPP + NOH), 1020 (M+, H2TFPP + N02), 1038 (M+, H2TFPP + HNO3), as well as peaks alluding to higher order addition of these substituents and their metal derivatives shed more lights on the nature and extent of these substitution reactions. Cu(II) or Ni (11) Adducts CuTFPP in HOAC-HNO3 produced one major green fraction, which was chromatographed on a silica gel plate with dichloromethane-hexane (95:5) and collected as the least polar band. The absorption maxima of this Cu ABSORPTION 126 4 I I I 5 I I I /\ 545 J 1/ ‘AOQL’ 400 500 600 nm Figure 4. Absorption spectrum of H Ag(4‘3) in CHzClz. 127 ,4 M K cm Figure 5a. Expanded B-pyrrolic region of 1H-NMR spectrum of Zn(Z-B), fraction 2, in c003. 128 L l I l J l l I J 38 33 3.0 as ea rl-m Figure Sb. Expanded B-pyrrolic region of 1i-I-NMR spectrum of Zn(Z-B), fraction 4, in CDC13. 129 (2-B) species, it demetallated derivative as well as those of CuTPP and its [3- nitro substituted analogs are given in Table 5. This table shows a systematic wavelength-shifts caused by the introduction of B-nitro group in TPP is reproduced in the TFPP adducts. More importantly, a FAB-MS peak at m/z 1081 (M+, CuTFPP+NOz) confirms the assignment. Table 5. Absorption Maxima (N m) of TFPP and TPP adducts in CH2C12. Compound Soret Visible Bands CuTFPP 408 533, 571 Cu(2-B) 416 543, 589 CuTPP 414 539 B-NOz-CuTPP 422 551 H2TFPP 412 506, 583, 655 ch(4-B) 419 518, 590, 646 HZTPP 417 514, 548, 592, 647 B-NOz-HZTPP 425 528, 561, 602, 664 NiTFPP reactivity is also analogous to that of Cu complex. One major fraction was obtained under the same condition described earlier for the copper derivative. Ni (2-B) species features absorption maxima at 423, 540, and 592 nm and a FAB-V15 peak at m/z 1076 (M+, NiTFPP + N02). Demetallation of Ni (2-B) in TFA resulted in the formation of HNi(4-B) with absorption maxima at 424, 520, 595 and 650 nm. This spectrum, with the exception of band I, is red-shifted relative to H2TFPP, analogous to that observed in HCu (4-B). 130 Ag(II) Adducts Demetallation of the reaction products of H2TFPP + AgNO3/HOAC in TFA followed by chromatography on alumina resulted in the separation of five green fractions. The absorption maxima and FAB-MS data are given in Table 6. Table 6. Absorption Maxima and FAB-Ms Data on AgTFPP Adducts. Elution No. Amax (nm) m/z 1 422, 550, 595 1097 (M+,AgTFPP+HzO) 2 428, 562, 604 1038 (M+,H2TFPP+HNO3) 3 419, 515, 556, 594, 657 1038 (M+,H2TFPP4-HN O3) 993 (M+,H2TFPP+HZO) 4 422, 516, 555, 594, 649 1038 (M+,H2TFPP+I-INO3) 5 418, 510, 552, 595, 644 993 (M+,H2TFPP-l-HZO) As it appears from the MS data in Table 6, addition of H20 and HNO3 moieties across the double bond and formation of chlorin-type structures is evident. Although these fractions have been purified on alumina column, they seem to be partially contaminated by one another. This is especially true for fraction three in which there seems to be equal amounts of water and nitric acid adducts present. If we rule out the possibility of functional group transformation, 11;. (Irolysis, tautomerism15 ..... during the acquisition of FAB— MS, then HPLC tel 'zmiques may be required to further purify these fractions. 131 Zn (II) Adducts Reaction“ of ZnTFPP with HOAC/HNO3 resulted in the separation of five major and a few minor fractions on silica plate after chromatography with dichloromethane-hexane (5%) as eluant. Table 7 summarizes absorption maxima and preliminary FAB-MS and 1H-NMR data for the major bands. The proton NMR spectra of the Zn adducts, exhibit a more defined picture of B-proton splittings than other derivatives and may be utilized as the subject of further investigation. The prominent MS peaks of the fractions 1 and 2 indicate a higher order of substitution with Zn complexes than the previous systems. The m/z at 1128 in fraction 1 corresponds to the addition of two nitro groups and m/z at 1173 in fraction 2 corresponds to the substitution of three nitro groups in the ZnTFPP complex. Table 7. Absorption Maxima (nm), FAB-M5 m/z and 1H-NMR B-proton Multiplicity Number (MN) of ZnTFPP Adducts. 1 436, 566, 613 1128 (MT) 2 (8.77 and 8.81 ppm) 2 435, 565, 610 1172 (MT) 9 or 10 1128 (M+-N02) 3 440, 572, 620 -- 3 (8,82, 8.86, 8.88 ppm) 4 439, 570-, 608 -- 9 5 436, 573, 619 -- multiple _ 132 Discussion The original reaction condition used by Gouterman and coworkers, i.e., H2TFPP+AgNO3/HOAC was intended to find a more effective method for silver ion insertion into H2TFPP. This is because the AgOAC/HOAC method also gives poor yield. Indeed, the AgNO3/HOAC condition facilitates the insertion of silver ion and metallation is quantitative in < 1 hour; however, the reaction condition immediately proceeds to form the green adducts. The FAB-M5 obtained on a sample which also refluxed for 3/ 4 hour already indicates the presence of addition products across Cbe double bond, Scheme 1. It is possible to speculate that the electron-deficient TFPP system triggers a H H H NOH H H H NO OH H H N/ N H N N ‘1 N /" H H + M+=993 M =1007 H H o H NO H No2 / L' N/ N H N N +— M+=1020 M -1038 Scheme 1 metal-catalyzed addition reaction to occur. Of the structures suggested, the M+ = 993 and 1038 peaks are more prominent in this sample than the others, Figure 6 a and b. ABSORPTION 4 I 9 I 537 320 I 570593. L L l L L I a 400 500 600 nm Figure 6a. 'l'he 593 nm band is due to the B-pyrrolic adducts. "Gouterman reports AgTFPP 133,, at 416, 536 and 570 nm in 41013. 134 £39..." 5: 64 3.3.3533 :5— .a a}: as... .39. — v 5 69:05 U‘ex' 363.8111?) CH,cu no») ll 0 ”a. >._._mzm...ZH 2424”. I600 |200 I400 FREQUENCY (cm") I000 200 400 600 800 Figure 5. RR spectra of CoOEPX . The solvent bands at 1422, 703 and 283 car1 are labeled with an asterisk. 149 2. According to Clark and Williams, the U(MBr)/0(MC1) and u(MI)/0(MC1) ratios are'0.77 - 0.74 and 0.65, respectively.18b 3. Nakagawa and Shimanouchi have obtained the infrared spectra of [Co(NH3)5X]2' and trans-[Co(NH3)4X2]+- type complexes.18C Table 2 lists the observed frequencies and band assignments obtained by these workers. 4. These workers also obtained the following force constants (mdyn/ A) for Co-N and Co-X bonds: K(Co-N), 1.05; K(Co-F), 0.99; K(Co-Cl), 0.91; K(Co- Br), 1.03,- and K(Co-I), 0.62.18d 5. In general, the infrared intensity of 0(MX) decreases in the order 0(MF)>1)(MC1)>1)(MBr)>u(MI), whereas the opposite order prevails for the Raman intensity.18C , Basically, there is very limited information available on the vibrational characteristics of five-coordinate cobaltic porphyrin complexes in the litera- ture. Beside the fact that 59Co natural abundance is nearly 100% and halide isotopic substitution is extremely difficult, Gouterman and Zerner have shown that the cobaltic CT transitions are mostly of (1r,d22) and (ddeZ) characterllrlz, in contrast with the ferric halide species in which X —) M CT transition in the near-UV region (i.e., 350 - 370 nm) is predominant.13 Thus, the RR spectra of C015 N-OEPX, Co-X derivatives of porphyrins other than OEP and IR studies of these systems will be necessary to assign this mode positively. Figure 6 depicts that FeOEPX Raman Spectra, excited at 363.8 nm in CHZCIZ, are also identical, however they do exhibit strong differences in the relative RR band intensities owing to the differences in the near-UV electronic transitions. The most striking difference is observed in the relative 150 >4” - 363.8 nm snap- snoutlll Fe(III)OEPX >._._mzm._.ZH Z<2._._.wszZH 24.24;. 200 400 600 800 FREQUENCY (cm") RR spectra of FeOEPX . Figure 6. 151 intensities of high and low frequency modes. A prominent feature is observed at 574 cm'1 in the fluoro adduct which is absent in all other derivatives. 54/56Fe frequency shifts (3 cm‘l) in CHzClz, Figure 7a, confirms the 574 cm'1 band assignment as Fe-F stretching vibration. This is possibly in contradiction to the previous assignment of 0(Fe-F) at 606 cm"1 in THF by Kitagawa and Kincaid”:15 The discrepancy, on the other hand, may arise from solvent effects, which are currently under investigation. 13d Interestingly, the well-studied 0(Fe-Cl) at ~ 360 cm- is not observed with excitation at 363.8 nm and in CH2C12, Figure 6. Figures 7b and 8, however, reveal the presence of this mode at 360 cm-1 with excitation at 406.7 and 413.1 nm in 0H. The 0(Fe—F), on the other hand, is not enhanced under this condition or it is obscured by the 6H peak at 606 crn-l, Figure 8. These observations lead to the notion that 0(Fe-X) are not only CT enhanced, as demonstrated by Hendrickson __t_ 81.13, but solvents may shift both these electronic transitions and the metal-ligand stretching frequency. The FeOEPBr spectrum features a mode at 274 cm'l, Figure 8, which is close to the reported value for 0(Fe-Br) at 270 cm-16 . Although there are other porphyrin vibrations in this region, as shown by the presence of a split mode at 245 and 270 cm'1 in the spectra of other derivatives, the single 274 cm'1 band of the bromo adduct most likely reflects the 0(Fe-Br) as well. 54F e substitution should confirm this assignment. The feature at 245 cm'1 in FeOEPI spectrum may not be due to 0(Fe-I) reported at 246 cm-16 owing to the presence of a similar mode at 246 cm'1 in the spectrum of the chloro derivative, Figure 8. The stretching force constants in the ferric halide complexes differ in the order K(Fe-F)> K(Fe-Cl)> K(Fe-Br), as reported by 152 >.:mzm:.ZH Z<2r_._m2m._.ZH Ztm2m._.ZH Z<§ M CT transitions are red-shifted or more likely vanished in the n-cation radicals. Since there are no differences in the absorption spectra of ferric porphyrin radicals, Figure 4, as there are in the corresponding neutral species, Figure 3, the CT transitions, based on uv-vis features, are expected not to be present or to be of very low probability. On the other hand, the cobaltic porphyrin radical for X = Br reveals a feature at ~ 160 cm'1 with 676.4 nm RR excitation (not shown) which we believe to be a mode involving metal and the bromide ligand.16 This mode is also observed in the spectrum excited at 406.7 nm (not shown), but not at 363.8 nm, Figure 9. The optical spectra of cobaltic-alkyl complexes differ substantially from those of six-coordinate Co(ROH)20EPX and five-coordinate CoOEPX: the absorption maxima of the Soret,[3 and 0: bands hardly shifts relative to those of CoOEP and remain at 390, 520 and 552 nm, respectively. The characteristic optical feature, however, is the nearly equal intensity of the a and [3 bands.9 CoOEPR (where R = Me, Et, i-Pr and t-Bu) displays an infrared mode at 343, 344, 339 and 343 cm‘l, respectively, analogous to that observed in methylcobalamine at 348 cm-1.17 The latter absorption was shown to be due to the 0(Co-C) by the slow disappearance of this band when the sample (KBr pallet) was irradiated with tungsten light: the photolysis product, hydroxocobalamine does not absorb at this frequency. The absence of any significant changes in the 0(Co-C) of these complexes may be the result of offsetting effects. The increase in the steric bulk of the alkyl chain, which tends to decrease the bond frequency by virtue of mass effect, is compensated 155 OVQNI ZBr' >._._mZMPZH 2325”. I000 |200 I400 I600 200 400 600 800 FREQUENCY (cm") Figure 9. RR spectra of ComOEP+°2X' i (a) X' = FeCl4' ~ C1”; (b) X' = Br‘. 156 by the greater electron-releasing ability of the more substituted alkyl group, which increases the force constant, bond order and bond frequency of the Co- C bond. Further confirmation of this mode requires isotopic substitution of the alkyl chain as well as preparation of Co-R derivatives of other porphyrins. Besides, a more complete investigation of 1)(CoC) measurement should also include other macrocyclic ligands such as chlorins and corrins to develop a better understanding of the use of the corrin ring in vitamin B12 prosthetic group, as a means of modulation of the Co-C bond strength. 10. 11. REFERENCES Buchler, I. W.; Kokisch, W.; Smith, P. D., Struct. Bonding 1978, 84, 79. Su, Y. 0.; Czernuszerwics, R. 5.; Miller, L. A.; Spiro, T. G., I. Am. Chem. Soc., 1988, 1_1_0_, 4150. (a) O'Keefe, D. H.; Barlow, C. H.; Smythe, G. A.; Fuchsman, W. H.; Moss, T. H.; Lilienthal, H. R; Caughy, W. 5., Bioinorg. Chem., 1975, 8, 125. (b) Lueken, H.; Buchler, I. W.; Lay, K. L.; Z. N aturforsh, 1976, 3_112, 1596. (c) Vaska, L.; Amundsen, a. R.; Brady, R.; Flynn, B. R.; Nakai, H., Finn. Chem. Lett., 1974, 66. (d) Ogoshi, H.; Watanabe, E.; Yoshida, Z.; Kincaid, I.; Nakamoto, K., I. Am. Chem. Soc., 1973, g, 2845. ' (e) Strauss, S. H.; Pawlik, M. I.; Skowyra, I.; Kennedy, I. R.; Anderson, 0. P.; Spartalian, K.; Dyne, I. L., Inorg. Chem. 1987, 26, 724. Hashimoto, 5.; Tatsuno, Y.; Kitagawa, T., I. Am. Chem. Soc., 1987,109, 8096. Geno, M. K.; Halpern, I., I. Am. Chem. Soc., 1987, 102, 1238. Ogoshi, H.; Watanabe, E.; Yoshida, Z.; Kincaid, I.; Nakamoto, K., I. Am. Chem. Soc., 1973, 22, 2845. Choi, 5.; Spiro, T. G.; I. Am. Chem. Soc., 1983, 186, 3683. Datta-Gupta, N .; Bardos, T. I., I. Pharm. Sci. 1968, 5_Z, 300. Ogoshi, H.; Watanabe, E. I.; Kokatsu, N .; Yoshida, Z. 1., Bull. Chem. Soc. Ipn., 1976, Q, 2529. Setsune, I. I.; Ikeda, M.; Kishimoto, Y.; Kitao, T.; I. Am. Chem. Soc., 1986, 188, 1309. (a) Tait, C. D.; Holten, D.; Gouterman, M., I. Am. Chem. Soc., 1984, 106, 6653. (b) Tait, C. D.; Holten, D.; gouterman, M., Chem. Phys. Lett., 1983, 100, 268. 157 12. 13. 14. 15. 16. 17. 18. 158 Edwards, W. D.; Zerner, M. 0, Can. I. Chem., 1985, 68, 1763. Hendrickson, D. N .; Kinnaird, M. G.; Suslick, K. 5., I. Am. Chem. Soc., 1987, 122, 1243. Kitagawa, T.; Abe, M.; Kyogoku, Y.; Ogoshi, H.; Watanabe, E.; Yoshida, Z., I. Phys. Chem., 1976, 8_0, 1181. Kincaid, I.; N akamoto, K., Spectros. Lett., 1976, 2, 19. Oertling, W. A.; Salehi, A.; Chung, Y. C.; Leroi, G. E.; Chang, C. K.; Babcock, G. T., I. Phys. Chem., _91, 5887. Hogenkamp, H. P. 0; Rush, I. E.; Swenson, C. A., I. Biol. Chem., 1965, Lg, 3641. (a) N akamoto, K., In "Infrared and Raman Spectra of Inorganic Coordination Compounds", Nakamoto, K., Ed., Iohn Wiley 8: Sons: New York, 1986, 324. (b) Clark, R I.; Williams, C. 5., Inorg. Chem., 1965, 4, 350. (c) Nakagawa, I.; Shimanouchi, T., Spectrochim. Acta, 1966, 2_2_, 759. (d) Shimanouchi, T.; Nakagawa, I.; Inorg. Chem., 1964, 8, 1805. (e) Ref. 18a, p. 379. HICHIGQN STRT E UNIV. LIBRARIES IIHWI Illl llllIlllllllllllllllllWIIHII 1 15792 3 2930078